This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

The optical geometry definition of the total deflection angle of a light ray in curved spacetime

Hideyoshi Arakida11footnotetext: Corresponding author.
Abstract

Assuming a static and spherically symmetric spacetime, we propose a novel concept of the total deflection angle of a light ray in terms of the optical geometry which is the Riemannian geometry experienced by the light ray. The total deflection angle is defined by the difference between the sum of internal angles of two triangles; one of the triangles lies on curved spacetime distorted by a gravitating body and the other on its background. The triangle required to define the total deflection angle can be realized by setting three laser-beam baselines as in planned space missions such as LATOR, ASTROD-GW, and LISA. Accordingly, the new total deflection angle is, in principle, measurable by gauging the internal angles of the triangles. The new definition of the total deflection angle can provide a geometrically and intuitively clear interpretation. Two formulas are proposed to calculate the total deflection angle on the basis of the Gauss–Bonnet theorem. It is shown that in the case of the Schwarzschild spacetime, the expression for the total deflection angle αSch\alpha_{\rm Sch} reduces to Epstein–Shapiro’s formula when the source of a light ray and the observer are located in an asymptotically flat region. Additionally, in the case of the Schwarzschild–de Sitter spacetime, the expression for the total deflection angle αSdS\alpha_{\rm SdS} comprises the Schwarzschild-like parts and coupling terms of the central mass mm and the cosmological constant Λ\Lambda in the form of 𝒪(Λm){\cal O}(\Lambda m) instead of 𝒪(Λ/m){\cal O}(\Lambda/m). Furthermore, αSdS\alpha_{\rm SdS} does not include the terms characterized only by the cosmological constant Λ\Lambda.

1 Introduction

Generally, the cosmological constant Λ\Lambda, or dark energy, is considered a promising candidate that can explain the accelerating expansion of the universe [1, 2, 3]. Although intensive studies have been performed for solving the mystery surrounding the cosmological constant/dark energy from both the theoretical and observational viewpoints, definitive direct evidence has not been obtained.

Because the structure of spacetime (the form of the metric gμνg_{\mu\nu}) depends on the existence of the cosmological constant Λ\Lambda, the evidence of the existence of the cosmological constant may be detected using the classical tests of general relativity, such as the observation of light deflection and the perihelion/periastron advance. Especially, light deflection is the basis of gravitational lensing which is a powerful tool in astrophysics and cosmology; for more details, see [4, 5] and the references therein. Therefore, gravitational lensing may prove the existence of the cosmological constant.

Thus far, the influence of the cosmological constant on light deflection has been studied in various ways, including whether the cosmological constant contributes to light deflection. Historically, Islam [6] first mentioned that light trajectory is independent of the cosmological constant Λ\Lambda because the second-order differential equation of the light ray does not include Λ\Lambda. Therefore, it was considered for a long time that Λ\Lambda did not contribute to light deflection. However, in 2007, Rindler and Ishak [7] indicated that Λ\Lambda affects the bending of a light ray, by using the invariant cosine formula under the Schwarzschild–de Sitter/Kottler solution. Starting with this paper [7], many authors intensively discussed its appearance in diverse ways; see [8] for a review article, and also see [9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20] and the references therein.

Despite intensive research in the past, a definitive conclusion has not been drawn, mainly because of the following reasons:

  • Unlike the Schwarzschild spacetime, the spacetime does not become asymptotically flat because of the cosmological constant Λ\Lambda. Accordingly, we cannot apply the standard procedure, which is described in many textbooks and literature, for calculating the total deflection angle.

  • [7] indicated that to calculate the angle in curved spacetime, one must focus on not only the equation of light trajectory but also the metric (ruler) of spacetime. However, in many studies, only the equation of light trajectory was considered when discussing the total deflection angle.

  • It is still not clear what is the total deflection angle and how it should be defined in curved spacetime.

Especially, the third reason above seems to be an essential problem that makes it difficult to clarify the contribution of the cosmological constant to total deflection angle. To overcome this difficulty, some authors applied the the Gauss–Bonnet theorem [21, 22, 23, 24, 25]. Although the Gauss–Bonnet theorem might help in solving the problem of the total deflection angle, it has still not been resolved and thus further consideration is necessary.

This study aims to provide a renewed method to solve the problem of total deflection angle and reveal the influence of the cosmological constant on light deflection. To this end, by assuming a static and spherically symmetric spacetime, we propose a new concept of the total deflection angle of a light ray in terms of the optical geometry which is regarded as the Riemannian geometry experienced by the light ray; the concept is realized by considering the difference between the sum of internal angles of two triangles; one exists in curved spacetime distorted by a gravitating body and the other in its background. This concept of the total deflection angle is inspired by space missions including LATOR [26], ASTROD-GW [27], and LISA [28], which set three laser-beams baselines in the space; accordingly, the new total deflection angle could be measured, in principle, by gauging the internal angle of each triangle. The new total deflection angle is geometrically and intuitively clear to interpret. To calculate the total deflection angle, we develop two formulas using the Gauss–Bonnet theorem.

This paper is outlined as follows. Section 2 introduces the optical metric that is considered the Riemannian geometry of light rays. Section 3 summarizes the Gauss–Bonnet theorem. Section 4 proposes a new concept and definition of the total deflection angle and presents two formulas to calculate the total deflection angle. In Sections 5 and 6, we apply the two formulas to compute the total deflection angle in the Schwarzschild and Schwarzschild–de Sitter spacetimes, respectively. Finally, conclusions are drawn in Section 7.

2 Optical Metric

We assume the following static and spherically symmetric spacetime: 222 It is possible to start from more generic spherical metric form as ds2=A(r)dt2+B(r)dr2+C(r)(dθ2+sin2θdϕ2)ds^{2}=-A(r)dt^{2}+B(r)dr^{2}+C(r)(d\theta^{2}+\sin^{2}\theta d\phi^{2}), and to construct the optical metric, see e.g., [22, 23, 25]. However, in this paper, we adopt a metric of type gtt=1/grrg_{tt}=-1/g_{rr} because we are interested in the propagation of light in Schwarzschild and Schwarzschild–de Sitter spacetimes.

ds2\displaystyle ds^{2} =gμνdxμdxν\displaystyle=g_{\mu\nu}dx^{\mu}dx^{\nu}
=f(r)dt2+1f(r)dr2+r2(dθ2+sin2θdϕ2),\displaystyle=-f(r)dt^{2}+\frac{1}{f(r)}dr^{2}+r^{2}(d\theta^{2}+\sin^{2}\theta d\phi^{2}), (2.1)

where gμνg_{\mu\nu} denotes the metric tensor of spacetime whose signature is denoted by ημν=diag(:+:+:+)\eta_{\mu\nu}={\rm diag}(-:+:+:+); f(r)f(r) denotes a function of the radial coordinate rr, Greek indices e.g., μ,ν\mu,\nu, range from 0 to 3, and we choose the geometrical unit c=G=1c=G=1 throughout this paper. Because of the spherical symmetry, we consider the equatorial plane θ=π/2,dθ=0\theta=\pi/2,d\theta=0 as the orbital plane of the light rays. One has

ds2=f(r)dt2+1f(r)dr2+r2dϕ2.\displaystyle ds^{2}=-f(r)dt^{2}+\frac{1}{f(r)}dr^{2}+r^{2}d\phi^{2}. (2.2)

Two constants of motion, energy EE and angular momentum LL, are expressed as

E=f(r)dtdλ,L=r2dϕdλ,\displaystyle E=f(r)\frac{dt}{d\lambda},\quad L=r^{2}\frac{d\phi}{d\lambda}, (2.3)

where λ\lambda denotes an affine parameter. Furthermore, the impact parameter bb is introduced as

bLE.\displaystyle b\equiv\frac{L}{E}. (2.4)

Using the null condition ds2=0ds^{2}=0, we introduce the optical metric g¯ij\bar{g}_{ij}, which is considered the Riemannian geometry experienced by light rays. One has

dt2\displaystyle dt^{2} g¯ijdxidxj=(gijg00)dxidxj\displaystyle\equiv\bar{g}_{ij}dx^{i}dx^{j}=\left(\frac{g_{ij}}{g_{00}}\right)dx^{i}dx^{j}
=g¯rrdr2+g¯ϕϕdϕ2\displaystyle=\bar{g}_{rr}dr^{2}+\bar{g}_{\phi\phi}d\phi^{2}
=1[f(r)]2dr2+r2f(r)dϕ2,\displaystyle=\frac{1}{[f(r)]^{2}}dr^{2}+\frac{r^{2}}{f(r)}d\phi^{2}, (2.5)

where Latin indices, e.g., i,ji,j, assume the values of i,j=1,2i,j=1,2, which correspond to 1=r1=r and 2=ϕ2=\phi, respectively. Hereafter, in accordance with [29], we refer to the geometry defined by the optical metric g¯ij\bar{g}_{ij} as the optical (reference) geometry opt{\cal M}^{\rm opt}. In the optical geometry description of Einstein’s relativity, light trajectories in the 3-dimensional space corresponding to a static spacetime are geodesic lines. This allows one to give a precise and practical definition of total deflection angle of the light ray.

Notably, on the optical reference geometry opt{\cal M}^{\rm opt}, it is observed from Eq. (2.5) that the time coordinate tt plays the role of an arc length parameter because

t1t2𝑑t=t1t2g¯rr(kr)2+g¯ϕϕ(kϕ)2𝑑t=t2t1,\displaystyle\int_{t_{1}}^{t_{2}}dt=\int_{t_{1}}^{t_{2}}\sqrt{\bar{g}_{rr}(k^{r})^{2}+\bar{g}_{\phi\phi}(k^{\phi})^{2}}dt=t_{2}-t_{1}, (2.6)

where ki=dxi/dtk^{i}=dx^{i}/dt denotes the unit tangent vector along the path of the light ray on opt{\cal M}^{\rm opt}, and it satisfies 1=g¯ijkikj1=\bar{g}_{ij}k^{i}k^{j} from Eq. (2.5). The property given in Eq. (2.6) is appropriate for application to the Gauss–Bonnet theorem later.

Let us summarize some important properties of the optical metric g¯μν\bar{g}_{\mu\nu}. If considering the slice of constant time tt of the spacetime Eq. (2.2), the spatial part of the metric gijg_{ij} is described as

d2\displaystyle d\ell^{2} gijdxidxj\displaystyle\equiv g_{ij}dx^{i}dx^{j}
=grrdr2+gϕϕdϕ2\displaystyle=g_{rr}dr^{2}+g_{\phi\phi}d\phi^{2}
=dr2f(r)+r2dϕ2.\displaystyle=\frac{dr^{2}}{f(r)}+r^{2}d\phi^{2}. (2.7)

Two metrics, Eqs. (2.5) and (2.7), are connected by the conformal transformation (conformal mapping) as

g¯ij=ω2(𝒙)gij,ω2(𝒙)=1f(r),\displaystyle\bar{g}_{ij}=\omega^{2}(\mbox{\boldmath$x$})g_{ij},\quad\omega^{2}(\mbox{\boldmath$x$})=\frac{1}{f(r)}, (2.8)

or more generally,

g¯μν=ω2(𝒙)gμν,\displaystyle\bar{g}_{\mu\nu}=\omega^{2}(\mbox{\boldmath$x$})g_{\mu\nu}, (2.9)

where ω2(𝒙)\omega^{2}(\mbox{\boldmath$x$}) denotes the conformal factor. Because the conformal transformation preserves the angle of the point at which the two curves intersect, the angles remain the same in both g¯ij\bar{g}_{ij} and gijg_{ij}. However, the conformal transformation rescales the coordinate value. Moreover, the null geodesic does not change its form upon performing the conformal transformation because of the null condition ds2=0ds^{2}=0; see Appendix G in [30].

3 Gauss–Bonnet Theorem

In the optical reference geometry opt{\cal M}^{\rm opt} given by the metric g¯ij\bar{g}_{ij}, (see Eq. (2.5)), we consider an nn-vertex polygon Σn\Sigma^{n}, which is orientable and bounded by nn smooth and piecewise regular curves Cp(p=1,2,,n)C_{p}~{}(p=1,2,\cdots,n) (see Figure 1). The (local) Gauss–Bonnet theorem is expressed as described in, e.g., on p. 139 in [31], p. 170 in [32], and p. 272 in [33], as follows:

ΣnK𝑑σ+p=1nCpκg𝑑t+p=1nθp=2π,\displaystyle\iint_{\Sigma^{n}}Kd\sigma+\sum_{p=1}^{n}\int_{C_{p}}\kappa_{g}dt+\sum_{p=1}^{n}\theta_{p}=2\pi, (3.1)

where an arc length parameter is denoted by tt instead of ss (see Eq. (2.6)), and an arc length parameter tt moves along the curve CpC_{p} in such a manner that a polygon Σn\Sigma^{n} stays on the left side; θp\theta_{p} denotes the external angle at the pp-th vertex, and θp\theta_{p} is determined as the sense leaving the internal angle on the left.

Refer to caption
Figure 1: Gauss–Bonnet theorem. A polygon Σn\Sigma^{n} is bounded by curves C1,C2,,CnC_{1},C_{2},\cdots,C_{n}, and the external angles of polygon Σn\Sigma^{n} are denoted by θp(p=1,2,,n)\theta_{p}~{}(p=1,2,\cdots,n).

KK denotes the Gaussian curvature as defined in, e.g., p. 147 of [32]

K=1g¯rrg¯ϕϕ[r(1g¯rrg¯ϕϕr)+ϕ(1g¯ϕϕg¯rrϕ)],\displaystyle K=-\frac{1}{\sqrt{\bar{g}_{rr}\bar{g}_{\phi\phi}}}\left[\frac{\partial}{\partial r}\left(\frac{1}{\sqrt{\bar{g}_{rr}}}\frac{\partial\sqrt{\bar{g}_{\phi\phi}}}{\partial r}\right)+\frac{\partial}{\partial\phi}\left(\frac{1}{\sqrt{\bar{g}_{\phi\phi}}}\frac{\partial\sqrt{\bar{g}_{rr}}}{\partial\phi}\right)\right], (3.2)

which represents the manner in which the spacetime is curved, and dσ=|det(g¯ij)|dx1dx2=|det(g¯ij)|drdϕd\sigma=\sqrt{|\det(\bar{g}_{ij})|}dx^{1}dx^{2}=\sqrt{|\det(\bar{g}_{ij})|}drd\phi denotes an areal element. The term κg\kappa_{g} denotes the geodesic curvature along curve CpC_{p}, e.g., on p. 256 in [33]

κg=12g¯rrg¯ϕϕ(g¯ϕϕrdϕdtg¯rrϕdrdt)+dΦdt,\displaystyle\kappa_{g}=\frac{1}{2\sqrt{\bar{g}_{rr}\bar{g}_{\phi\phi}}}\left(\frac{\partial\bar{g}_{\phi\phi}}{\partial r}\frac{d\phi}{dt}-\frac{\partial\bar{g}_{rr}}{\partial\phi}\frac{dr}{dt}\right)+\frac{d\Phi}{dt}, (3.3)

where Φ\Phi denotes the angle between the radial unit vector erie_{r}^{i} along radial geodesics and the tangent vector ki=dxi/dtk^{i}=dx^{i}/dt of curve CpC_{p}. The term κg\kappa_{g} characterizes the extent to which curve CpC_{p} deviates from the geodesic. Accordingly, if curve CpC_{p} is the geodesic, κg=0\kappa_{g}=0.

4 New Concept of the Total Deflection Angle

We propose a new concept of the total deflection angle α\alpha. Intuitively, the total deflection angle of a light ray is the change in the direction of light ray in the presence and absence of a gravitating body. However, the metric of spacetime is different in the presence and absence of gravitating body, thus each null geodesic essentially exists in a distinct spacetime determined by a different metric (ruler); for instance, in the case of the Schwarzschild spacetime, one null geodesic lies on curved spacetime (curved ruler), while the another exists in flat spacetime (flat ruler). Therefore, it becomes difficult to compare two null geodesics with each other in the same spacetime (same ruler). As an exception, the total deflection angle can be obtained only when both observer RR and light source SS are placed in asymptotically flat regions, owing to the Euclidean parallel postulate, which enables us to determine the angle at a distant point from the angle of any point PP, such as the corresponding angle and the alternate angle. Notably, the total deflection angle in the Schwarzschild spacetime can be obtained as the twice of angle ψP\psi_{P} at PP, i.e., α=2ψP\alpha=2\psi_{P}, where PP denotes the light source SS or observer RR. However, in a curved spacetime or region, the parallel postulate does not hold, and thus α2ψP\alpha\neq 2\psi_{P}.

However, even in curved spacetime, the internal angles of polygon Σn\Sigma^{n} can be measured by using the equipment mounted on the spacecrafts located at each vertex. Accordingly, the sum of the internal angles of polygon Σn\Sigma^{n} can be calculated. Because the sum of the internal angles of the polygon depends on the curvature of the spacetime, one might consider the difference between the sum of internal angles of two polygons that placed in distinct spacetimes. Therefore, we define the renewed total deflection angle α\alpha as the difference between the sums of the internal angles of two polygons.

To feasibly realize the above-mentioned concept of total deflection angle, we construct triangle Σ3\Sigma^{3} on the optical reference geometry opt{\cal M}^{\rm opt}; the triangle is bounded by three null geodesics Γ1\Gamma_{1}, Γ2\Gamma_{2}, and Γ3\Gamma_{3}. Practically, these three null geodesics can be substantiated by three laser-beam baselines that connect three spacecrafts, or two spacecrafts and International Space Station/ground station on the Earth, as in planned missions, e.g., LATOR, ASTROD-GW, and LISA. For more details, see Figure 2, wherein RR, MM, and SS denote the triangle vertexes where the satellites or ISS/ground stations are located. We denote the impact parameters of three null geodesics Γ1\Gamma_{1}, Γ2\Gamma_{2}, and Γ3\Gamma_{3} by b1b_{1}, b2b_{2}, and b3b_{3}, respectively. For simplicity, we arrange triangle Σ3\Sigma^{3} such that the angular coordinates ϕ\phi at the closest approach of Γ1\Gamma_{1}, Γ2\Gamma_{2}, and Γ3\Gamma_{3} correspond to ϕ=π/2\phi=\pi/2, ϕ=π/2δ2\phi=\pi/2-\delta_{2}, and ϕ=π/2+δ3\phi=\pi/2+\delta_{3}, respectively. ΓR\Gamma_{R}, ΓM\Gamma_{M}, and ΓS\Gamma_{S} correspond to the coordinate lines ϕ=ϕR\phi=\phi_{R}, ϕ=ϕM\phi=\phi_{M}, and ϕ=ϕS\phi=\phi_{S}, respectively passing through the points RR, MM, and SS, respectively. ΓR\Gamma_{R}, ΓM\Gamma_{M}, and ΓS\Gamma_{S} are related to the unit radial vector in the equatorial plane, noting that we are working in the optical reference geometry opt{\cal M}^{\rm opt},

eri=(1g¯rr,0).\displaystyle e^{i}_{r}=\left(\frac{1}{\sqrt{\bar{g}_{rr}}},0\right). (4.1)

We note that for the observer at the points RR, MM, and SS, ΓR\Gamma_{R}, ΓM\Gamma_{M}, and ΓS\Gamma_{S} can be regarded as the radial null geodesics of the light rays coming from the central object OO 333 It may be noteworthy that in terms of the null geodesics, Γ1\Gamma_{1}, Γ2\Gamma_{2}, Γ3\Gamma_{3}, ΓR\Gamma_{R}, ΓM\Gamma_{M}, and ΓS\Gamma_{S} are more precisely spatial geodesics which are the projections of null geodesics on the spatial spacetime sections t=constt={\rm const}. However, due to the nature of conformal transformation, the form of the differential equation of the photon trajectory dr/dϕdr/d\phi on the optical metric is the same as the form of the equation on the spatial spacetime sections t=constt={\rm const} derived from the null condition ds2=0ds^{2}=0., however, from the point of view of the coordinate system, ΓR\Gamma_{R}, ΓM\Gamma_{M}, and ΓS\Gamma_{S} are not radial because the observer cannot be at r=0r=0.

Refer to caption
Figure 2: Configuration of triangle Σ3\Sigma^{3}. A triangle Σ3\Sigma^{3} is bounded by three null geodesics, Γ1\Gamma_{1}, Γ2\Gamma_{2}, and Γ3\Gamma 3, whose impact parameters are denoted by b1b_{1}, b2b_{2}, and b3b_{3}, respectively. RR, MM, and SS denote the triangle vertexes where the satellites or ISS/ground stations are located. For simplicity, the three null geodesics are arranged such that the point of the closest approach corresponds to ϕ=π/2\phi=\pi/2, ϕ=π/2δ2\phi=\pi/2-\delta_{2}, and ϕ=π/2+δ3\phi=\pi/2+\delta_{3}, respectively. ΓR\Gamma_{R}, ΓM\Gamma_{M}, and ΓS\Gamma_{S} correspond to the coordinate lines ϕ=ϕR\phi=\phi_{R}, ϕ=ϕM\phi=\phi_{M}, and ϕ=ϕS\phi=\phi_{S}, respectively passing through the points RR, MM, and SS, respectively.

Let us define the renewed total deflection angle in accordance with the above-mentioned configuration of triangle Σ3\Sigma^{3}. Because Γ1\Gamma_{1}, Γ2\Gamma_{2}, and Γ3\Gamma_{3} are null geodesics, the line integral of geodesic curvature κg\kappa_{g} vanishes. One has

Cpκg𝑑t=Γpκg𝑑t=0,p=1,2,3.\displaystyle\int_{C_{p}}\kappa_{g}dt=\int_{\Gamma_{p}}\kappa_{g}dt=0,\quad p=1,2,3. (4.2)

Accordingly, the Gauss–Bonnet theorem, Eq. (3.1), reduces to

Σ3K𝑑σ+p=13θp=2π.\displaystyle\iint_{\Sigma^{3}}Kd\sigma+\sum_{p=1}^{3}\theta_{p}=2\pi. (4.3)

Next, we prepare the same relation as that in Eq. (4.3) for the background spacetime, e.g., Minkowski and de Sitter spacetimes, and write the following:

Σ3KBG𝑑σBG+p=13θpBG=2π,\displaystyle\iint_{\Sigma^{3}}K^{\rm BG}d\sigma^{\rm BG}+\sum_{p=1}^{3}\theta^{\rm BG}_{p}=2\pi, (4.4)

where superscript BG{\rm BG} denotes the background. Additionally, for the same reason as that in the case of Eq. (4.3), the line integral of geodesic curvature κgBG\kappa_{g}^{\rm BG} is also zero.

Rewriting the sum of the external angles, θp\theta_{p}, using the internal angles βp\beta_{p}, we have

p=13θp\displaystyle\sum_{p=1}^{3}\theta_{p} =3πp=13βp,\displaystyle=3\pi-\sum_{p=1}^{3}\beta_{p}, (4.5)
p=13θpBG\displaystyle\sum_{p=1}^{3}\theta_{p}^{\rm BG} =3πp=13βpBG,\displaystyle=3\pi-\sum_{p=1}^{3}\beta^{\rm BG}_{p}, (4.6)

where we used the relation θp=πβp\theta_{p}=\pi-\beta_{p}. Subtracting Eq. (4.4) from Eq. (4.3) and using Eqs. (4.5) and (4.6), we obtain the following relation:

Σ3K𝑑σΣ3KBG𝑑σBG=p=13(βpβpBG).\displaystyle\iint_{\Sigma^{3}}Kd\sigma-\ \iint_{\Sigma^{3}}K^{\rm BG}d\sigma^{\rm BG}=\sum_{p=1}^{3}\left(\beta_{p}-\beta^{\rm BG}_{p}\right). (4.7)

Using the right-hand side of Eq. (4.7), let us define the renewed total deflection angle as

α|p=13(βpβpBG)|.\displaystyle\alpha\equiv\left|\sum_{p=1}^{3}\left(\beta_{p}-\beta^{\rm BG}_{p}\right)\right|. (4.8)

Eq. (4.8) represents the difference in the sum of internal angles βp\beta_{p} between two triangles. Eq. (4.8) provides an instinctive and clear definition of the total deflection angle. The absolute-value symbol in Eq. (4.8) indicates that we take the total deflection angle to be a positive value.

To obtain the internal angles βp\beta_{p}, we compute ψp\psi_{p}, which denote the intersection angles between the null geodesics (Γp(p=1,2,3))(\Gamma_{p}~{}~{}(p=1,2,3)) and radial null geodesics (ΓR(\Gamma_{R} ΓM\Gamma_{M}, and ΓS)\Gamma_{S}). Angles ψp\psi_{p} can be calculated using the tangent formula as follows:

tanψp=g¯ϕϕ(rp)g¯rr(rp)dϕdrp=f(rp)rpdϕdrp,p=1,2,3.\displaystyle\tan\psi_{p}=\frac{\sqrt{\bar{g}_{\phi\phi}(r_{p})}}{\sqrt{\bar{g}_{rr}(r_{p})}}\frac{d\phi}{dr_{p}}=\sqrt{f(r_{p})}r_{p}\frac{d\phi}{dr_{p}},\quad p=1,2,3. (4.9)

Using ψp\psi_{p}, the internal angle can be calculated as

βR\displaystyle\beta_{R} =ψ2(ϕR)ψ1(ϕR),\displaystyle=\psi_{2}(\phi_{R})-\psi_{1}(\phi_{R}), (4.10)
βM\displaystyle\beta_{M} =ψ3(ϕM)ψ2(ϕM)+π,\displaystyle=\psi_{3}(\phi_{M})-\psi_{2}(\phi_{M})+\pi, (4.11)
βS\displaystyle\beta_{S} =ψ1(ϕS)ψ3(ϕS),\displaystyle=\psi_{1}(\phi_{S})-\psi_{3}(\phi_{S}), (4.12)

where the subscripts, 11, 22, and 33, of ψ\psi represent the intersection angle between light trajectory, Γ1\Gamma_{1}, Γ2\Gamma_{2}, and Γ3\Gamma_{3}, and radial null geodesics. Additionally, ϕS\phi_{S}, ϕM\phi_{M}, and ϕR\phi_{R} denote the angular coordinate values at points SS, MM, and RR, respectively. See Figure 2. Notably, angle ψp\psi_{p} is determined to be in the counterclockwise direction from the light trajectory, Γ1\Gamma_{1}, Γ2\Gamma_{2}, and Γ3\Gamma_{3} to the radial geodesics, ΓR\Gamma_{R}, ΓM\Gamma_{M}, and ΓS\Gamma_{S}, (see Figure 2). Notably, Eq. (4.9) gives angle ψp\psi_{p} by considering the metric, namely, the part f(rp)\sqrt{f(r_{p})}, in Eq. (4.9).

Using Eq. (4.7), the renewed total deflection angle α\alpha can be also expressed as the difference in the areal integral of the Gaussian curvature KK as

α|Σ3K𝑑σΣ3KBG𝑑σBG|.\displaystyle\alpha\equiv\left|\iint_{\Sigma^{3}}Kd\sigma-\iint_{\Sigma^{3}}K^{\rm BG}d\sigma^{\rm BG}\right|. (4.13)

We refer to Eq. (4.8) as the angular formula of the total deflection angle and Eq. (4.13) the integral formula of the total deflection angle.

5 Total Deflection Angle in the Schwarzschild Spacetime

We first examine the total deflection angle in the Schwarzschild spacetime as

fSch(r)=12mr,\displaystyle f^{\rm Sch}(r)=1-\frac{2m}{r}, (5.1)

where mm denotes the mass of the central (lens) object.

5.1 Light Trajectory

The first-order differential equation for null geodesic is given as

(drpSchdϕ)2=(rpSch)2[(rpSch)2bp21+2mrpSch],p=1,2,3,\displaystyle\left(\frac{dr^{\rm Sch}_{p}}{d\phi}\right)^{2}=(r^{\rm Sch}_{p})^{2}\left[\frac{(r^{\rm Sch}_{p})^{2}}{b^{2}_{p}}-1+\frac{2m}{r^{\rm Sch}_{p}}\right],\quad p=1,2,3, (5.2)

where bpb_{p} denotes an impact parameter of null geodesic Γp\Gamma_{p}. Changing the variable upSch=1/rpSchu^{\rm Sch}_{p}=1/r^{\rm Sch}_{p}, Eq. (5.2) becomes

(dupSchdϕ)2=1bp2(upSch)2+2m(upSch)3.\displaystyle\left(\frac{du^{\rm Sch}_{p}}{d\phi}\right)^{2}=\frac{1}{b^{2}_{p}}-(u^{\rm Sch}_{p})^{2}+2m(u^{\rm Sch}_{p})^{3}. (5.3)

We derive the trajectories of null geodesics Γ1\Gamma_{1}, Γ2\Gamma_{2}, and Γ3\Gamma_{3}, three of which configure the triangle on the optical reference geometry opt{\cal M}^{\rm opt} (see Figure 2). In these cases, the zeroth-order solutions are

u1Sch,0=sinϕb1,u2Sch,0=sin(ϕ+δ2)b2,u3Sch,0=sin(ϕδ3)b3.\displaystyle u^{\rm Sch,0}_{1}=\frac{\sin\phi}{b_{1}},\quad u^{\rm Sch,0}_{2}=\frac{\sin(\phi+\delta_{2})}{b_{2}},\quad u^{\rm Sch,0}_{3}=\frac{\sin(\phi-\delta_{3})}{b_{3}}. (5.4)

In accordance with the standard perturbation scheme, we express the solution upSch=upSch(ϕ)u^{\rm Sch}_{p}=u^{\rm Sch}_{p}(\phi) as

upSch=upSch,0+εupSch,1+ε2upSch,2,\displaystyle u^{\rm Sch}_{p}=u^{\rm Sch,0}_{p}+\varepsilon u^{\rm Sch,1}_{p}+\varepsilon^{2}u^{\rm Sch,2}_{p}, (5.5)

where ε\varepsilon denotes the small dimensionless expansion parameter, which in the Schwarzschild spacetime is ε=m/b\varepsilon=m/b. The terms εupSch,1\varepsilon u^{\rm Sch,1}_{p} and ε2upSch,2\varepsilon^{2}u^{\rm Sch,2}_{p} denote the first 𝒪(ε){\cal O}(\varepsilon) and second 𝒪(ε2){\cal O}(\varepsilon^{2}) order correction terms, respectively. Substituting Eq (5.5) into Eq. (5.3), we obtain the second-order solution with respect to ε\varepsilon as

u1Sch\displaystyle u^{\rm Sch}_{1} =sinϕb1+m2b12(3+cos2ϕ)\displaystyle=\frac{\sin\phi}{b_{1}}+\frac{m}{2b_{1}^{2}}(3+\cos 2\phi)
+m216b13[37sinϕ+30(π2ϕ)cosϕ3sin3ϕ]+𝒪(ε3),\displaystyle+\frac{m^{2}}{16b_{1}^{3}}\left[37\sin\phi+30(\pi-2\phi)\cos\phi-3\sin 3\phi\right]+{\cal O}(\varepsilon^{3}), (5.6)
u2Sch\displaystyle u^{\rm Sch}_{2} =sin(ϕ+δ2)b2+m2b22[3+cos2(ϕ+δ2)]\displaystyle=\frac{\sin(\phi+\delta_{2})}{b_{2}}+\frac{m}{2b_{2}^{2}}\left[3+\cos 2(\phi+\delta_{2})\right]
+m216b23{37sin(ϕ+δ2)+30[π2(ϕ+δ2)]cos(ϕ+δ2)3sin3(ϕ+δ2)}+𝒪(ε3),\displaystyle+\frac{m^{2}}{16b_{2}^{3}}\left\{37\sin(\phi+\delta_{2})+30[\pi-2(\phi+\delta_{2})]\cos(\phi+\delta_{2})-3\sin 3(\phi+\delta_{2})\right\}+{\cal O}(\varepsilon^{3}), (5.7)
u3Sch\displaystyle u^{\rm Sch}_{3} =sin(ϕδ3)b3+m2b32[3+cos2(ϕδ3)]\displaystyle=\frac{\sin(\phi-\delta_{3})}{b_{3}}+\frac{m}{2b_{3}^{2}}\left[3+\cos 2(\phi-\delta_{3})\right]
+m216b33{37sin(ϕδ3)+30[π2(ϕδ3)]cos(ϕδ3)3sin3(ϕδ3)}+𝒪(ε3).\displaystyle+\frac{m^{2}}{16b_{3}^{3}}\left\{37\sin(\phi-\delta_{3})+30[\pi-2(\phi-\delta_{3})]\cos(\phi-\delta_{3})-3\sin 3(\phi-\delta_{3})\right\}+{\cal O}(\varepsilon^{3}). (5.8)

The integration constants of Eqs. (5.6), (5.7), and (5.8) are chosen to maximize uu (or minimize rr) at ϕ=π/2\phi=\pi/2, ϕ=π/2δ2\phi=\pi/2-\delta_{2}, and ϕ=π/2+δ3\phi=\pi/2+\delta_{3}, respectively:

du1Schdϕ|ϕ=π/2=0,du2Schdϕ|ϕ=π/2δ2=0,du3Schdϕ|ϕ=π/2+δ3=0.\displaystyle\left.\frac{du^{\rm Sch}_{1}}{d\phi}\right|_{\phi=\pi/2}=0,\quad\left.\frac{du^{\rm Sch}_{2}}{d\phi}\right|_{\phi=\pi/2-\delta_{2}}=0,\quad\left.\frac{du^{\rm Sch}_{3}}{d\phi}\right|_{\phi=\pi/2+\delta_{3}}=0. (5.9)

5.2 Angular Formula

We compute the total deflection angle αSch\alpha_{\rm Sch} by using the angular formula, i.e., Eq. (4.8). First, we calculate the intersection angles ψ\psi between the null geodesics Γ1\Gamma_{1}, Γ2\Gamma_{2}, and Γ3\Gamma_{3}, and the radial null geodesics, ΓR\Gamma_{R}, ΓM\Gamma_{M}, and ΓS\Gamma_{S}, respectively. Substituting Eqs. (5.2), (5.6), (5.7), and (5.8) into Eq. (4.9) and expanding up to the second-order with respect to ε\varepsilon, we have

ψ1Sch\displaystyle\psi_{1}^{\rm Sch} =ϕ+2mb1cosϕ+m28b12[15(π2ϕ)sin2ϕ]+𝒪(ε3),\displaystyle=\phi+\frac{2m}{b_{1}}\cos\phi+\frac{m^{2}}{8b_{1}^{2}}\left[15(\pi-2\phi)-\sin 2\phi\right]+{\cal O}(\varepsilon^{3}), (5.10)
ψ2Sch\displaystyle\psi_{2}^{\rm Sch} =ϕ+δ2+2mb2cos(ϕ+δ2)+m28b22[15(π2ϕ2δ2)sin2(ϕ+δ2)]+𝒪(ε3),\displaystyle=\phi+\delta_{2}+\frac{2m}{b_{2}}\cos(\phi+\delta_{2})+\frac{m^{2}}{8b_{2}^{2}}\left[15(\pi-2\phi-2\delta_{2})-\sin 2(\phi+\delta_{2})\right]+{\cal O}(\varepsilon^{3}), (5.11)
ψ3Sch\displaystyle\psi_{3}^{\rm Sch} =ϕδ3+2mb3cos(ϕδ3)+m28b32[15(π2ϕ+2δ3)sin2(ϕδ3)]+𝒪(ε3).\displaystyle=\phi-\delta_{3}+\frac{2m}{b_{3}}\cos(\phi-\delta_{3})+\frac{m^{2}}{8b_{3}^{2}}\left[15(\pi-2\phi+2\delta_{3})-\sin 2(\phi-\delta_{3})\right]+{\cal O}(\varepsilon^{3}). (5.12)

Using Eqs. (4.10), (4.11), and (4.12), internal angles βR\beta_{R}, βM\beta_{M}, and βS\beta_{S} are given as

βRSch\displaystyle\beta_{R}^{\rm Sch} =δ2+2mb2cos(ϕR+δ2)2mb1cosϕR\displaystyle=\delta_{2}+\frac{2m}{b_{2}}\cos(\phi_{R}+\delta_{2})-\frac{2m}{b_{1}}\cos\phi_{R}
+m28b22[15(π2ϕR2δ2)sin2(ϕR+δ2)]\displaystyle+\frac{m^{2}}{8b_{2}^{2}}\left[15(\pi-2\phi_{R}-2\delta_{2})-\sin 2(\phi_{R}+\delta_{2})\right]
m28b12[15(π2ϕR)sin2ϕR]+𝒪(ε3),\displaystyle-\frac{m^{2}}{8b_{1}^{2}}\left[15(\pi-2\phi_{R})-\sin 2\phi_{R}\right]+{\cal O}(\varepsilon^{3}), (5.13)
βMSch\displaystyle\beta_{M}^{\rm Sch} =δ3δ2+π+2mb3cos(ϕMδ3)2mb2cos(ϕM+δ2)\displaystyle=\delta_{3}-\delta_{2}+\pi+\frac{2m}{b_{3}}\cos(\phi_{M}-\delta_{3})-\frac{2m}{b_{2}}\cos(\phi_{M}+\delta_{2})
+m28b32[15(π2ϕM+2δ3)sin2(ϕMδ3)]\displaystyle+\frac{m^{2}}{8b_{3}^{2}}\left[15(\pi-2\phi_{M}+2\delta_{3})-\sin 2(\phi_{M}-\delta_{3})\right]
m28b22[15(π2ϕM2δ2)sin2(ϕM+δ2)]+𝒪(ε3),\displaystyle-\frac{m^{2}}{8b_{2}^{2}}\left[15(\pi-2\phi_{M}-2\delta_{2})-\sin 2(\phi_{M}+\delta_{2})\right]+{\cal O}(\varepsilon^{3}), (5.14)
βSSch\displaystyle\beta_{S}^{\rm Sch} =δ3+2mb1cosϕS2mb3cos(ϕSδ3)\displaystyle=-\delta_{3}+\frac{2m}{b_{1}}\cos\phi_{S}-\frac{2m}{b_{3}}\cos(\phi_{S}-\delta_{3})
+m28b12[15(π2ϕS)sin2ϕS]\displaystyle+\frac{m^{2}}{8b_{1}^{2}}\left[15(\pi-2\phi_{S})-\sin 2\phi_{S}\right]
m28b32[15(π2ϕS+2δ3)sin2(ϕSδ3)]+𝒪(ε3).\displaystyle-\frac{m^{2}}{8b_{3}^{2}}\left[15(\pi-2\phi_{S}+2\delta_{3})-\sin 2(\phi_{S}-\delta_{3})\right]+{\cal O}(\varepsilon^{3}). (5.15)

Because the background of the Schwarzschild spacetime is the flat Minkowski spacetime, one has

p=13βpBG=p=13βpMin=π.\displaystyle\sum_{p=1}^{3}\beta^{\rm BG}_{p}=\sum_{p=1}^{3}\beta^{\rm Min}_{p}=\pi. (5.16)

Substituting Eqs. (5.13), (5.14), (5.15), and (5.16) into (4.8), we obtain the total deflection angle as

αSch\displaystyle\alpha_{\rm Sch} =|p=13(βpSchβpMin)|=π(βRSch+βMSch+βSSch)\displaystyle=\left|\sum_{p=1}^{3}(\beta^{\rm Sch}_{p}-\beta^{\rm Min}_{p})\right|=\pi-(\beta^{\rm Sch}_{R}+\beta^{\rm Sch}_{M}+\beta^{\rm Sch}_{S})
=2m[cosϕRcosϕSb1+cos(ϕM+δ2)cos(ϕR+δ2)b2+cos(ϕSδ3)cos(ϕMδ3)b3]\displaystyle=2m\left[\frac{\cos\phi_{R}-\cos\phi_{S}}{b_{1}}+\frac{\cos(\phi_{M}+\delta_{2})-\cos(\phi_{R}+\delta_{2})}{b_{2}}+\frac{\cos(\phi_{S}-\delta_{3})-\cos(\phi_{M}-\delta_{3})}{b_{3}}\right]
m24[sin2ϕRsin2ϕS2b12+sin2(ϕM+δ2)sin2(ϕR+δ2)2b22\displaystyle-\frac{m^{2}}{4}\left[\frac{\sin 2\phi_{R}-\sin 2\phi_{S}}{2b_{1}^{2}}+\frac{\sin 2(\phi_{M}+\delta_{2})-\sin 2(\phi_{R}+\delta_{2})}{2b_{2}^{2}}\right.
+sin2(ϕSδ3)sin2(ϕMδ3)2b3215(ϕRϕSb12+ϕMϕRb22+ϕSϕMb32)]\displaystyle+\frac{\sin 2(\phi_{S}-\delta_{3})-\sin 2(\phi_{M}-\delta_{3})}{2b_{3}^{2}}-\left.15\left(\frac{\phi_{R}-\phi_{S}}{b_{1}^{2}}+\frac{\phi_{M}-\phi_{R}}{b_{2}^{2}}+\frac{\phi_{S}-\phi_{M}}{b_{3}^{2}}\right)\right]
+𝒪(ε3),\displaystyle+{\cal O}(\varepsilon^{3}), (5.17)

where the sign of αSch\alpha_{\rm Sch} is taken to be positive. Notably, the internal angles βp\beta_{p} can be measured, in principle, via actual observations using a spacecraft.

5.3 Integral Formula

We show that Eq. (5.17) can be also obtained using the integral formula, i.e., Eq. (4.13). On the optical metric Eq. (2.5), the Gaussian curvature, i.e., Eq. (3.2), is given as

KSch=2mr3(13m2r)<0,\displaystyle K^{\rm Sch}=-\frac{2m}{r^{3}}\left(1-\frac{3m}{2r}\right)<0, (5.18)

and the areal element dσSchd\sigma^{\rm Sch} becomes

dσSch=r(12mr)32drdϕ.\displaystyle d\sigma^{\rm Sch}=r\left(1-\frac{2m}{r}\right)^{-\frac{3}{2}}drd\phi. (5.19)

Because the background of the Schwarzschild spacetime is the flat Minkowski spacetime, the Gaussian curvature is KMin=0K^{\rm Min}=0, and

Σ3KBG𝑑σBG=Σ3KMin𝑑σMin=0.\displaystyle\iint_{\Sigma^{3}}K^{\rm BG}d\sigma^{\rm BG}=\iint_{\Sigma^{3}}K^{\rm Min}d\sigma^{\rm Min}=0. (5.20)

We divide triangle Σ3\Sigma^{3}, which is bounded by three geodesics Γ1\Gamma_{1}, Γ2\Gamma_{2}, and Γ3\Gamma_{3}, into two parts, namely, ΣRM3(ϕRϕϕM)\Sigma^{3}_{RM}(\phi_{R}\leq\phi\leq\phi_{M}) and ΣMS3(ϕMϕϕS)\Sigma^{3}_{MS}(\phi_{M}\leq\phi\leq\phi_{S}), by assuming ϕR<ϕM<ϕS\phi_{R}<\phi_{M}<\phi_{S} (see Figure 3).

Refer to caption
Figure 3: Dividing triangle Σ3\Sigma^{3} into two parts. To integrate the areal integral of the Gaussian curvature KK over the triangle Σ3\Sigma^{3}, we divide Σ3\Sigma^{3} into ΣRM3(ϕRϕϕM)\Sigma^{3}_{RM}(\phi_{R}\leq\phi\leq\phi_{M}) and ΣMS3(ϕMϕϕS)\Sigma^{3}_{MS}(\phi_{M}\leq\phi\leq\phi_{S}). Here we assume that the angular coordinates satisfy the magnitude relation, i.e., ϕR<ϕM<ϕS\phi_{R}<\phi_{M}<\phi_{S}.

Expanding up to the second order with respect to ε=m/b\varepsilon=m/b, the areal integral of KSchK^{\rm Sch} yields the total deflection angle as

αSch\displaystyle\alpha_{\rm Sch} =|Σ3KSch𝑑σSchΣ3KMin𝑑σMin|=Σ3KSch𝑑σSch\displaystyle=\left|\iint_{\Sigma^{3}}K^{\rm Sch}d\sigma^{\rm Sch}-\iint_{\Sigma^{3}}K^{\rm Min}d\sigma^{\rm Min}\right|=-\iint_{\Sigma^{3}}K^{\rm Sch}d\sigma^{\rm Sch}
=ϕRϕMr1Schr2Sch(2mr2+3m2r3)𝑑r𝑑ϕ+ϕMϕSr1Schr3Sch(2mr2+3m2r3)𝑑r𝑑ϕ+𝒪(ε3)\displaystyle=\int_{\phi_{R}}^{\phi_{M}}\int_{r^{\rm Sch}_{1}}^{r^{\rm Sch}_{2}}\left(\frac{2m}{r^{2}}+\frac{3m^{2}}{r^{3}}\right)drd\phi+\int_{\phi_{M}}^{\phi_{S}}\int_{r^{\rm Sch}_{1}}^{r^{\rm Sch}_{3}}\left(\frac{2m}{r^{2}}+\frac{3m^{2}}{r^{3}}\right)drd\phi+{\cal O}(\varepsilon^{3})
=2m[cosϕRcosϕSb1+cos(ϕM+δ2)cos(ϕR+δ2)b2+cos(ϕSδ3)cos(ϕMδ3)b3]\displaystyle=2m\left[\frac{\cos\phi_{R}-\cos\phi_{S}}{b_{1}}+\frac{\cos(\phi_{M}+\delta_{2})-\cos(\phi_{R}+\delta_{2})}{b_{2}}+\frac{\cos(\phi_{S}-\delta_{3})-\cos(\phi_{M}-\delta_{3})}{b_{3}}\right]
m24[sin2ϕRsin2ϕS2b12+sin2(ϕM+δ2)sin2(ϕR+δ2)2b22\displaystyle-\frac{m^{2}}{4}\left[\frac{\sin 2\phi_{R}-\sin 2\phi_{S}}{2b_{1}^{2}}+\frac{\sin 2(\phi_{M}+\delta_{2})-\sin 2(\phi_{R}+\delta_{2})}{2b_{2}^{2}}\right.
+sin2(ϕSδ3)sin2(ϕMδ3)2b3215(ϕRϕSb12+ϕMϕRb22+ϕSϕMb32)]\displaystyle+\frac{\sin 2(\phi_{S}-\delta_{3})-\sin 2(\phi_{M}-\delta_{3})}{2b_{3}^{2}}-\left.15\left(\frac{\phi_{R}-\phi_{S}}{b_{1}^{2}}+\frac{\phi_{M}-\phi_{R}}{b_{2}^{2}}+\frac{\phi_{S}-\phi_{M}}{b_{3}^{2}}\right)\right]
+𝒪(ε3),\displaystyle+{\cal O}(\varepsilon^{3}), (5.21)

where we take the sign of αSch\alpha_{\rm Sch} to be positive. Additionally, r1Schr^{\rm Sch}_{1}, r2Schr^{\rm Sch}_{2}, and r3Schr^{\rm Sch}_{3} are given by Eqs. (5.6), (5.7), and (5.8), respectively.

5.4 Limit of Infinite Source-Observer Distance

Let us confirm that Eqs. (5.17) and (5.21) can reproduce Epstein–Shapiro’s formula [34]. To this end, as shown in Figure 4, we re-arrange the triangle such that it is symmetric with respect to ϕ=π/2\phi=\pi/2; we put ϕM=π/2\phi_{M}=\pi/2, b2=b3b_{2}=b_{3}, and δ3=δ2\delta_{3}=\delta_{2}. Additionally, letting source SS and observer RR to be located at asymptotically infinite flat regions ϕSπ\phi_{S}\rightarrow\pi and ϕR0\phi_{R}\rightarrow 0, respectively, we obtain

αSch4m(1b1sinδ2+cosδ2b2)+m24(15πb12+2sin2δ215πb22)+𝒪(ε3),\displaystyle\alpha^{\rm Sch}\rightarrow 4m\left(\frac{1}{b_{1}}-\frac{\sin\delta_{2}+\cos\delta_{2}}{b_{2}}\right)+\frac{m^{2}}{4}\left(\frac{15\pi}{b_{1}^{2}}+\frac{2\sin 2\delta_{2}-15\pi}{b^{2}_{2}}\right)+{\cal O}(\varepsilon^{3}), (5.22)

where the following two terms:

4msinδ2+cosδ2b2,m242sin2δ215πb22,\displaystyle-4m\frac{\sin\delta_{2}+\cos\delta_{2}}{b_{2}},\quad\frac{m^{2}}{4}\frac{2\sin 2\delta_{2}-15\pi}{b^{2}_{2}},

are newly appeared. However, as the two points SS and RR approach infinity, the impact parameter b2b_{2} of Γ2\Gamma_{2} and Γ3\Gamma_{3} becomes infinite, i.e., b2b_{2}\rightarrow\infty (see Figure 5).

Refer to caption
Figure 4: Symmetric triangle. We re-arrange triangle Σ3\Sigma^{3} such that it is symmetrical with respect to ϕ=π/2\phi=\pi/2, and we set b2=b3b_{2}=b_{3} and δ2=δ3\delta_{2}=\delta_{3}. The source SS and observer RR are located at infinity.

Accordingly, Eq. (5.22) results in

αSch4mb1+15πm24b12+𝒪(ε3).\displaystyle\alpha^{\rm Sch}\rightarrow\frac{4m}{b_{1}}+\frac{15\pi m^{2}}{4b_{1}^{2}}+{\cal O}(\varepsilon^{3}). (5.23)

Notably, null geodesics Γ2\Gamma_{2} and Γ3\Gamma_{3} are not the asymptotes of null geodesic Γ1\Gamma_{1}.

Refer to caption
Figure 5: Relationship between two points SS and RR of triangle Σ3\Sigma^{3} and its size. This figure shows that as the two points SS and RR approach infinity, the impact parameter b2b_{2} of Γ2\Gamma_{2} and Γ3\Gamma_{3} also becomes infinite, i.e., b2b_{2}\rightarrow\infty.

6 Total Deflection Angle in the Schwarzschild–de Sitter Spacetime

Let us derive the expression for the total deflection angle and investigate the contribution of the cosmological constant Λ\Lambda to the total deflection angle in the Schwarzschild–de Sitter/Kottler spacetime [35], which is characterized as

fSdS(r)=12mrΛ3r2,\displaystyle f^{\rm SdS}(r)=1-\frac{2m}{r}-\frac{\Lambda}{3}r^{2}, (6.1)

where Λ\Lambda denotes the cosmological constant.

6.1 Light Trajectory

The first-order differential equation for null geodesics becomes

(drpSdSdϕ)2=(rpSdS)2[(rpSdS)2(1bp2+Λ3)1+2mrpSdS].\displaystyle\left(\frac{dr^{\rm SdS}_{p}}{d\phi}\right)^{2}=(r^{\rm SdS}_{p})^{2}\left[(r^{\rm SdS}_{p})^{2}\left(\frac{1}{b^{2}_{p}}+\frac{\Lambda}{3}\right)-1+\frac{2m}{r^{\rm SdS}_{p}}\right]. (6.2)

Here, let us introduce another constant BB as

1Bp21bp2+Λ3,\displaystyle\frac{1}{B^{2}_{p}}\equiv\frac{1}{b^{2}_{p}}+\frac{\Lambda}{3}, (6.3)

and we rewrite Eq. (6.2) as

(drpSdSdϕ)2=(rpSdS)2[(rpSdS)2Bp21+2mrpSdS],\displaystyle\left(\frac{dr^{\rm SdS}_{p}}{d\phi}\right)^{2}=(r^{\rm SdS}_{p})^{2}\left[\frac{(r^{\rm SdS}_{p})^{2}}{B^{2}_{p}}-1+\frac{2m}{r^{\rm SdS}_{p}}\right], (6.4)

which is of the same form as Eq. (5.2). Therefore, up to the second order in ε=m/B\varepsilon=m/B instead of ε=m/b\varepsilon=m/b, the equation of the light trajectory in the Schwarzschild–de Sitter spacetime can be expressed as of the same form of Eqs. (5.6), (5.7), and (5.8),

u1SdS\displaystyle u_{1}^{\rm SdS} =sinϕB1+m2B12(3+cos2ϕ)\displaystyle=\frac{\sin\phi}{B_{1}}+\frac{m}{2B_{1}^{2}}(3+\cos 2\phi)
+m216B13[37sinϕ+30(π2ϕ)cosϕ3sin3ϕ]+𝒪(ε3),\displaystyle+\frac{m^{2}}{16B_{1}^{3}}\left[37\sin\phi+30(\pi-2\phi)\cos\phi-3\sin 3\phi\right]+{\cal O}(\varepsilon^{3}), (6.5)
u2SdS\displaystyle u_{2}^{\rm SdS} =sin(ϕ+δ2)B2+m2B22[3+cos2(ϕ+δ2)]\displaystyle=\frac{\sin(\phi+\delta_{2})}{B_{2}}+\frac{m}{2B_{2}^{2}}\left[3+\cos 2(\phi+\delta_{2})\right]
+m216B23{37sin(ϕ+δ2)+30[π2(ϕ+δ2)]cos(ϕ+δ2)3sin3(ϕ+δ2)}+𝒪(ε3),\displaystyle+\frac{m^{2}}{16B_{2}^{3}}\left\{37\sin(\phi+\delta_{2})+30[\pi-2(\phi+\delta_{2})]\cos(\phi+\delta_{2})-3\sin 3(\phi+\delta_{2})\right\}+{\cal O}(\varepsilon^{3}), (6.6)
u3SdS\displaystyle u_{3}^{\rm SdS} =sin(ϕδ3)B3+m2B32[3+cos2(ϕδ3)]\displaystyle=\frac{\sin(\phi-\delta_{3})}{B_{3}}+\frac{m}{2B_{3}^{2}}\left[3+\cos 2(\phi-\delta_{3})\right]
+m216B33{37sin(ϕδ3)+30[π2(ϕδ3)]cos(ϕδ3)3sin3(ϕδ3)}+𝒪(ε3).\displaystyle+\frac{m^{2}}{16B_{3}^{3}}\left\{37\sin(\phi-\delta_{3})+30[\pi-2(\phi-\delta_{3})]\cos(\phi-\delta_{3})-3\sin 3(\phi-\delta_{3})\right\}+{\cal O}(\varepsilon^{3}). (6.7)

The equation of light trajectory in the Schwarzschild–de Sitter spacetime does not depend on the cosmological constant Λ\Lambda and impact parameter bb. Using Eq. (6.4) and the condition

drdϕ|r=r0=0,\displaystyle\left.\frac{dr}{d\phi}\right|_{r=r_{0}}=0, (6.8)

we have following relation:

1B2=1b2+Λ3=1r022mr03,\displaystyle\frac{1}{B^{2}}=\frac{1}{b^{2}}+\frac{\Lambda}{3}=\frac{1}{r_{0}^{2}}-\frac{2m}{r_{0}^{3}}, (6.9)

where r0r_{0} denotes the radial coordinate value at the closest approach of the light ray. In principle, r0r_{0} can be obtained via actual measurements, as the circumference radius 0=2πr0\ell_{0}=2\pi r_{0}. Therefore, BB is calculated without knowing the value of Λ\Lambda and bb. However, as will be discussed below, this does not mean that the total deflection angle is also independent of Λ\Lambda.

6.2 Background of the Schwarzschild–de Sitter Spacetime

Before discussing the total deflection angle, let us debate the background of the Schwarzschild–de Sitter spacetime and derive some relations.

In the case of the Schwarzschild–de Sitter spacetime, the background should be regarded as the de Sitter spacetime instead of the Minkowski spacetime. This is because we had, in advance, assumed the existence of non-zero cosmological constant Λ\Lambda, which is not an integration constant as mass mm. In fact, the action

𝒮=[c416πG(R2Λ)+M]gd4x\displaystyle{\cal S}=\int\left[\frac{c^{4}}{16\pi G}(R-2\Lambda)+{\cal L}_{M}\right]\sqrt{-g}d^{4}x (6.10)

and the field equation

Rμν12gμνR+Λgμν=8πGc4Tμν\displaystyle R_{\mu\nu}-\frac{1}{2}g_{\mu\nu}R+\Lambda g_{\mu\nu}=\frac{8\pi G}{c^{4}}T_{\mu\nu} (6.11)

explicitly include the cosmological constant Λ\Lambda, where g=det(gμν)g=\det(g_{\mu\nu}); the term M{\cal L}_{M} denotes the Lagrangian for the matter field; RμνR_{\mu\nu} and RR denote the Ricci tensor and Ricci scalar, respectively; TμνT_{\mu\nu} denotes the energy-momentum tensor. Hence, the spacetime cannot reduce to the Minkowski spacetime.

The de Sitter spacetime is characterized by

fdS(r)=1Λ3r2,\displaystyle f^{\rm dS}(r)=1-\frac{\Lambda}{3}r^{2}, (6.12)

and the differential equation of a light ray on the optical reference geometry opt{\cal M}^{\rm opt} becomes

(drpdSdϕ)2=(rpdS)2[(1bp2+Λ3)(rpdS)21].\displaystyle\left(\frac{dr^{\rm dS}_{p}}{d\phi}\right)^{2}=(r^{\rm dS}_{p})^{2}\left[\left(\frac{1}{b^{2}_{p}}+\frac{\Lambda}{3}\right)(r^{\rm dS}_{p})^{2}-1\right]. (6.13)

Using Eq. (6.3), the light trajectories in the de Sitter spacetime are

u1dS\displaystyle u_{1}^{\rm dS} =sinϕB1,\displaystyle=\frac{\sin\phi}{B_{1}}, (6.14)
u2dS\displaystyle u_{2}^{\rm dS} =sin(ϕ+δ2)B2,\displaystyle=\frac{\sin(\phi+\delta_{2})}{B_{2}}, (6.15)
u3dS\displaystyle u_{3}^{\rm dS} =sin(ϕδ3)B3.\displaystyle=\frac{\sin(\phi-\delta_{3})}{B_{3}}. (6.16)

Substituting Eqs. (6.12), (6.14), (6.15), and (6.16) into Eq. (4.9), the intersection angles ψ\psi between three null geodesics Γp(p=1,2,3)\Gamma_{p}~{}(p=1,2,3) and radial null geodesics ΓR\Gamma_{R}, ΓM\Gamma_{M}, and ΓS\Gamma_{S}, respectively, are computed as

ψ1dS\displaystyle\psi^{\rm dS}_{1} =arctan(1ΛB123csc2ϕtanϕ)=arccoscosϕ1ΛB123,\displaystyle=\arctan\left(\sqrt{1-\frac{\Lambda B^{2}_{1}}{3}\csc^{2}\phi}\tan\phi\right)=\arccos\frac{\cos\phi}{\sqrt{1-\frac{\Lambda B_{1}^{2}}{3}}}, (6.17)
ψ2dS\displaystyle\psi^{\rm dS}_{2} =arctan(1ΛB223csc2(ϕ+δ2)tan(ϕ+δ2))=arccoscos(ϕ+δ2)1ΛB223,\displaystyle=\arctan\left(\sqrt{1-\frac{\Lambda B^{2}_{2}}{3}\csc^{2}(\phi+\delta_{2})}\tan(\phi+\delta_{2})\right)=\arccos\frac{\cos(\phi+\delta_{2})}{\sqrt{1-\frac{\Lambda B_{2}^{2}}{3}}}, (6.18)
ψ3dS\displaystyle\psi^{\rm dS}_{3} =arctan(1ΛB323csc2(ϕδ3)tan(ϕδ3))=arccoscos(ϕδ3)1ΛB323,\displaystyle=\arctan\left(\sqrt{1-\frac{\Lambda B^{2}_{3}}{3}\csc^{2}(\phi-\delta_{3})}\tan(\phi-\delta_{3})\right)=\arccos\frac{\cos(\phi-\delta_{3})}{\sqrt{1-\frac{\Lambda B_{3}^{2}}{3}}}, (6.19)

where we used the following inverse trigonometric function:

arctanϕ=arccos11+ϕ2.\displaystyle\arctan\phi=\arccos\frac{1}{\sqrt{1+\phi^{2}}}. (6.20)

Using Eqs. (4.10), (4.11), and (4.12), the sum of the internal angles, βpdS\beta^{\rm dS}_{p}, is given by

p=13βpdS\displaystyle\sum_{p=1}^{3}\beta^{\rm dS}_{p} =π\displaystyle=\pi
+arccoscosϕR1ΛB123arccoscosϕS1ΛB123\displaystyle+\arccos\frac{\cos\phi_{R}}{\sqrt{1-\frac{\Lambda B^{2}_{1}}{3}}}-\arccos\frac{\cos\phi_{S}}{\sqrt{1-\frac{\Lambda B^{2}_{1}}{3}}}
+arccoscos(ϕM+δ2)1ΛB223arccoscos(ϕR+δ2)1ΛB223\displaystyle+\arccos\frac{\cos(\phi_{M}+\delta_{2})}{\sqrt{1-\frac{\Lambda B^{2}_{2}}{3}}}-\arccos\frac{\cos(\phi_{R}+\delta_{2})}{\sqrt{1-\frac{\Lambda B^{2}_{2}}{3}}}
+arccoscos(ϕSδ3)1ΛB323arccoscos(ϕMδ3)1ΛB323.\displaystyle+\arccos\frac{\cos(\phi_{S}-\delta_{3})}{\sqrt{1-\frac{\Lambda B^{2}_{3}}{3}}}-\arccos\frac{\cos(\phi_{M}-\delta_{3})}{\sqrt{1-\frac{\Lambda B^{2}_{3}}{3}}}. (6.21)

Notably, the sum of the internal angles of the triangle differs from π\pi because of the existence of the cosmological constant Λ\Lambda.

On the other hand, the Gaussian curvature and areal element are given by

KdS=Λ3<0,dσdS=r(1Λ3r2)32drdϕ.\displaystyle K^{\rm dS}=-\frac{\Lambda}{3}<0,\quad d\sigma^{\rm dS}=r\left(1-\frac{\Lambda}{3}r^{2}\right)^{-\frac{3}{2}}drd\phi. (6.22)

As in the case of the Schwarzschild spacetime, we divide triangle Σ3\Sigma^{3} into ΣRM3(ϕRϕϕM)\Sigma^{3}_{RM}(\phi_{R}\leq\phi\leq\phi_{M}) and ΣMS3(ϕMϕϕS)\Sigma^{3}_{MS}(\phi_{M}\leq\phi\leq\phi_{S}) (see Figure 3). Subsequently, using Eq. (6.22), the areal integral of KdSK^{\rm dS} becomes

Σ3KdS𝑑σdS\displaystyle-\iint_{\Sigma^{3}}K^{\rm dS}d\sigma^{\rm dS} =ϕRϕMr1dSr2dSΛ3r(1Λ3r2)32𝑑r𝑑ϕ+ϕMϕSr1dSr3dSΛ3r(1Λ3r2)32𝑑r𝑑ϕ\displaystyle=\int^{\phi_{M}}_{\phi_{R}}\int^{r_{2}^{\rm dS}}_{r_{1}^{\rm dS}}\frac{\Lambda}{3}r\left(1-\frac{\Lambda}{3}r^{2}\right)^{-\frac{3}{2}}drd\phi+\int^{\phi_{S}}_{\phi_{M}}\int^{r_{3}^{\rm dS}}_{r_{1}^{\rm dS}}\frac{\Lambda}{3}r\left(1-\frac{\Lambda}{3}r^{2}\right)^{-\frac{3}{2}}drd\phi
=ϕRϕM[sin(ϕ+δ2)sin2(ϕ+δ2)ΛB223sinϕsin2ϕΛB123]𝑑ϕ\displaystyle=\int^{\phi_{M}}_{\phi_{R}}\left[\frac{\sin(\phi+\delta_{2})}{\sqrt{\sin^{2}(\phi+\delta_{2})-\frac{\Lambda B^{2}_{2}}{3}}}-\frac{\sin\phi}{\sqrt{\sin^{2}\phi-\frac{\Lambda B^{2}_{1}}{3}}}\right]d\phi
+ϕMϕS[sin(ϕδ3)sin2(ϕδ3)ΛB323sinϕsin2ϕΛB123]𝑑ϕ\displaystyle+\int^{\phi_{S}}_{\phi_{M}}\left[\frac{\sin(\phi-\delta_{3})}{\sqrt{\sin^{2}(\phi-\delta_{3})-\frac{\Lambda B^{2}_{3}}{3}}}-\frac{\sin\phi}{\sqrt{\sin^{2}\phi-\frac{\Lambda B^{2}_{1}}{3}}}\right]d\phi
=arccoscosϕS1ΛB123arccoscosϕR1ΛB123\displaystyle=\arccos\frac{\cos\phi_{S}}{\sqrt{1-\frac{\Lambda B^{2}_{1}}{3}}}-\arccos\frac{\cos\phi_{R}}{\sqrt{1-\frac{\Lambda B^{2}_{1}}{3}}}
+arccoscos(ϕR+δ2)1ΛB223arccoscos(ϕM+δ2)1ΛB223\displaystyle+\arccos\frac{\cos(\phi_{R}+\delta_{2})}{\sqrt{1-\frac{\Lambda B^{2}_{2}}{3}}}-\arccos\frac{\cos(\phi_{M}+\delta_{2})}{\sqrt{1-\frac{\Lambda B^{2}_{2}}{3}}}
+arccoscos(ϕMδ3)1ΛB323,arccoscos(ϕSδ3)1ΛB323\displaystyle+\arccos\frac{\cos(\phi_{M}-\delta_{3})}{\sqrt{1-\frac{\Lambda B^{2}_{3}}{3}}},-\arccos\frac{\cos(\phi_{S}-\delta_{3})}{\sqrt{1-\frac{\Lambda B^{2}_{3}}{3}}} (6.23)

where we used the following pseudo-elliptic integral, which is provided on p. 205 in [36];

sinϕa2sin2ϕ1𝑑ϕ=1aarcsinacosϕa21,a2>1,\displaystyle\int\frac{\sin\phi}{\sqrt{a^{2}\sin^{2}\phi-1}}d\phi=-\frac{1}{a}\arcsin\frac{a\cos\phi}{\sqrt{a^{2}-1}},\quad a^{2}>1, (6.24)

and the inverse trigonometric function: arcsinϕ+arccosϕ=π/2\arcsin\phi+\arccos\phi=\pi/2. Eqs. (6.21) and (6.23) are calculated without any approximation with respect to Λ\Lambda.

Before concluding this subsection, we give some comments. The sum of the internal angles of the triangle in the de Sitter spacetime differs from π\pi because of the non-zero Gaussian curvature KdS=Λ/3K^{\rm dS}=-\Lambda/3. Accordingly, the value of the cosmological constant Λ\Lambda may be obtained by taking the difference between p=13βpdS\sum_{p=1}^{3}\beta^{\rm dS}_{p} and π\pi. However, as indicated previously, if we consider the existence of non-zero cosmological constant in the context of the Schwarzschild–de Sitter spacetime, then the de Sitter spacetime should be adopted as the background instead of the Minkowski spacetime. In fact, the Schwarzschild–de Sitter spacetime can be considered the distorted de Sitter spacetime because of mass mm (see section 14.4 in [37])

6.3 Angular Formula

First, we obtain angle ψ\psi. We wish to express ψSdS\psi^{\rm SdS} as

ψpSdS=ψpdS+𝒪(m,m2,mΛ)terms,\displaystyle\psi_{p}^{\rm SdS}=\psi_{p}^{\rm dS}+{\cal O}(m,m^{2},m\Lambda)~{}\mbox{terms}, (6.25)

where ψpdS(p=1,2,3)\psi_{p}^{\rm dS}~{}(p=1,2,3) are given by Eqs. (6.17), (6.18), and (6.19), respectively.

To calculate the correction term 𝒪(m,m2,mΛ){\cal O}(m,m^{2},m\Lambda), we first set the small dimensionless expansion parameter ε\varepsilon as ε=m/B\varepsilon=m/B. Subsequently, using Eqs. (6.1), (6.4), (6.5), (6.6), and (6.7) we expand Eq. (4.9) up to the second order in ε=m/B\varepsilon=m/B and obtain the following approximate expression:

ψ1SdS\displaystyle\psi^{\rm SdS}_{1} =arccoscosϕ1ΛB123+2mcosϕB11ΛB123csc2ϕ\displaystyle=\arccos\frac{\cos\phi}{\sqrt{1-\frac{\Lambda B_{1}^{2}}{3}}}+\frac{2m\cos\phi}{B_{1}\sqrt{1-\frac{\Lambda B^{2}_{1}}{3}\csc^{2}\phi}}
+m216B12(1ΛB123csc2ϕ)32{2[15(π2ϕ)sin2ϕ]\displaystyle+\frac{m^{2}}{16B^{2}_{1}\left(1-\frac{\Lambda B^{2}_{1}}{3}\csc^{2}\phi\right)^{\frac{3}{2}}}\{2[15(\pi-2\phi)-\sin 2\phi]
ΛB12csc2ϕ(30π60ϕ+23cotϕ+9cos3ϕcscϕ)}+𝒪(ε3),\displaystyle-\Lambda B^{2}_{1}\csc^{2}\phi(30\pi-60\phi+23\cot\phi+9\cos 3\phi\csc\phi)\}+{\cal O}(\varepsilon^{3}), (6.26)
ψ2SdS\displaystyle\psi^{\rm SdS}_{2} =arccoscos(ϕ+δ2)1ΛB223+2mcos(ϕ+δ2)B21ΛB223csc2(ϕ+δ2)\displaystyle=\arccos\frac{\cos(\phi+\delta_{2})}{\sqrt{1-\frac{\Lambda B_{2}^{2}}{3}}}+\frac{2m\cos(\phi+\delta_{2})}{B_{2}\sqrt{1-\frac{\Lambda B^{2}_{2}}{3}\csc^{2}(\phi+\delta_{2})}}
+m216B22[1ΛB223csc2(ϕ+δ2)]32(2{15[π2(ϕ+δ2)]sin2(ϕ+δ2)}\displaystyle+\frac{m^{2}}{16B^{2}_{2}\left[1-\frac{\Lambda B^{2}_{2}}{3}\csc^{2}(\phi+\delta_{2})\right]^{\frac{3}{2}}}\left(2\{15[\pi-2(\phi+\delta_{2})]-\sin 2(\phi+\delta_{2})\}\right.
ΛB22csc2(ϕ+δ2)[30π60(ϕ+δ2)\displaystyle-\left.\Lambda B^{2}_{2}\csc^{2}(\phi+\delta_{2})[30\pi-60(\phi+\delta_{2})\right.
+23cot(ϕ+δ2)+9cos3(ϕ+δ2)csc(ϕ+δ2)])+𝒪(ε3),\displaystyle+\left.23\cot(\phi+\delta_{2})+9\cos 3(\phi+\delta_{2})\csc(\phi+\delta_{2})]\right)+{\cal O}(\varepsilon^{3}), (6.27)
ψ3SdS\displaystyle\psi^{\rm SdS}_{3} =arccoscos(ϕδ3)1ΛB323+2mcos(ϕδ3)B31ΛB323csc2(ϕδ3)\displaystyle=\arccos\frac{\cos(\phi-\delta_{3})}{\sqrt{1-\frac{\Lambda B_{3}^{2}}{3}}}+\frac{2m\cos(\phi-\delta_{3})}{B_{3}\sqrt{1-\frac{\Lambda B^{2}_{3}}{3}\csc^{2}(\phi-\delta_{3})}}
+m216B32[1ΛB323csc2(ϕδ3)]32(2{15[π2(ϕδ3)]sin2(ϕδ3)}\displaystyle+\frac{m^{2}}{16B^{2}_{3}\left[1-\frac{\Lambda B^{2}_{3}}{3}\csc^{2}(\phi-\delta_{3})\right]^{\frac{3}{2}}}\left(2\{15[\pi-2(\phi-\delta_{3})]-\sin 2(\phi-\delta_{3})\}\right.
ΛB32csc2(ϕδ3)[30π60(ϕδ3)\displaystyle-\left.\Lambda B^{2}_{3}\csc^{2}(\phi-\delta_{3})[30\pi-60(\phi-\delta_{3})\right.
+23cot(ϕδ3)+9cos3(ϕδ3)csc(ϕδ3)])+𝒪(ε3).\displaystyle+\left.23\cot(\phi-\delta_{3})+9\cos 3(\phi-\delta_{3})\csc(\phi-\delta_{3})]\right)+{\cal O}(\varepsilon^{3}). (6.28)

Note that at this stage, we do not adopt ε=ΛB2\varepsilon=\Lambda B^{2} as the small dimensionless expansion parameter. The first terms in the first lines of Eqs. (6.26), (6.27), and (6.28) are equal to ψ1dS\psi^{\rm dS}_{1}, ψ2dS\psi^{\rm dS}_{2}, and ψ3dS\psi^{\rm dS}_{3}, respectively (see Eqs. (6.17), (6.18), and (6.19)).

Next, expanding 𝒪(m){\cal O}(m) and 𝒪(m2){\cal O}(m^{2}) terms in Eqs. (6.26), (6.27), and (6.28) with respect to ε=ΛB2\varepsilon=\Lambda B^{2} and the remaining 𝒪(m/B,(m/B)2,(m/B)ΛB2){\cal O}(m/B,(m/B)^{2},(m/B)\cdot\Lambda B^{2}) terms, we have

ψ1SdS\displaystyle\psi^{\rm SdS}_{1} =ψ1dS+2mB1cosϕ+m28B12[15(π2ϕ)sin2ϕ]\displaystyle=\psi^{\rm dS}_{1}+\frac{2m}{B_{1}}\cos\phi+\frac{m^{2}}{8B_{1}^{2}}[15(\pi-2\phi)-\sin 2\phi]
+ΛmB13cotϕcscϕ+𝒪(ε3),\displaystyle+\frac{\Lambda mB_{1}}{3}\cot\phi\csc\phi+{\cal O}(\varepsilon^{3}), (6.29)
ψ2SdS\displaystyle\psi^{\rm SdS}_{2} =ψ2dS+2mB2cos2(ϕ+δ2)+m28B22{15[π2(ϕ+δ2)]sin2(ϕ+δ2)}\displaystyle=\psi^{\rm dS}_{2}+\frac{2m}{B_{2}}\cos 2(\phi+\delta_{2})+\frac{m^{2}}{8B_{2}^{2}}\{15[\pi-2(\phi+\delta_{2})]-\sin 2(\phi+\delta_{2})\}
+ΛmB23cot(ϕ+δ2)csc(ϕ+δ2)+𝒪(ε3),\displaystyle+\frac{\Lambda mB_{2}}{3}\cot(\phi+\delta_{2})\csc(\phi+\delta_{2})+{\cal O}(\varepsilon^{3}), (6.30)
ψ3SdS\displaystyle\psi^{\rm SdS}_{3} =ψ3dS+2mB3cos(ϕδ3)+m28B32{15[π2(ϕδ3)]sin2(ϕδ3)}\displaystyle=\psi^{\rm dS}_{3}+\frac{2m}{B_{3}}\cos(\phi-\delta_{3})+\frac{m^{2}}{8B_{3}^{2}}\{15[\pi-2(\phi-\delta_{3})]-\sin 2(\phi-\delta_{3})\}
+ΛmB33cot(ϕδ3)csc(ϕδ3)+𝒪(ε3),\displaystyle+\frac{\Lambda mB_{3}}{3}\cot(\phi-\delta_{3})\csc(\phi-\delta_{3})+{\cal O}(\varepsilon^{3}), (6.31)

where the residual terms 𝒪(ε3){\cal O}(\varepsilon^{3}) are 𝒪((m/B)3,(m/B)2ΛB2,(m/B)(ΛB2)2){\cal O}((m/B)^{3},(m/B)^{2}\cdot\Lambda B^{2},(m/B)\cdot(\Lambda B^{2})^{2}). Notably, ψ1SdS\psi^{\rm SdS}_{1}, ψ2SdS\psi^{\rm SdS}_{2}, and ψ3SdS\psi^{\rm SdS}_{3} comprise the terms characterized by the cosmological constant Λ\Lambda, namely ψ1dS\psi^{\rm dS}_{1}, ψ2dS\psi^{\rm dS}_{2}, and ψ3dS\psi^{\rm dS}_{3}. However as we will see below, the terms ψ1dS\psi^{\rm dS}_{1}, ψ2dS\psi^{\rm dS}_{2}, and ψ3dS\psi^{\rm dS}_{3} are eliminated from the expression of the total deflection angle αSdS\alpha_{\rm SdS}.

Using Eqs. (4.8), (4.10), (4.11), (4.12), (6.21), (6.29), (6.30), and (6.31), the total deflection angle becomes

αSdS\displaystyle\alpha_{\rm SdS} =|p=13(βpSdSβpdS)|=(βRdS+βMdS+βSdS)(βRSdS+βMSdS+βSSdS)\displaystyle=\left|\sum_{p=1}^{3}(\beta^{\rm SdS}_{p}-\beta^{\rm dS}_{p})\right|=(\beta_{R}^{\rm dS}+\beta_{M}^{\rm dS}+\beta_{S}^{\rm dS})-(\beta_{R}^{\rm SdS}+\beta_{M}^{\rm SdS}+\beta_{S}^{\rm SdS})
=2m[cosϕRcosϕSB1+cos(ϕM+δ2)cos(ϕR+δ2)B2+cos(ϕSδ3)cos(ϕMδ3)B3]\displaystyle=2m\left[\frac{\cos\phi_{R}-\cos\phi_{S}}{B_{1}}+\frac{\cos(\phi_{M}+\delta_{2})-\cos(\phi_{R}+\delta_{2})}{B_{2}}+\frac{\cos(\phi_{S}-\delta_{3})-\cos(\phi_{M}-\delta_{3})}{B_{3}}\right]
m24[sin2ϕRsin2ϕS2B12+sin2(ϕM+δ2)sin2(ϕR+δ2)2B22\displaystyle-\frac{m^{2}}{4}\left[\frac{\sin 2\phi_{R}-\sin 2\phi_{S}}{2B_{1}^{2}}+\frac{\sin 2(\phi_{M}+\delta_{2})-\sin 2(\phi_{R}+\delta_{2})}{2B_{2}^{2}}\right.
sin2(ϕMδ3)sin2(ϕSδ3)2B3215(ϕRϕSB12+ϕMϕRB22ϕMϕSB32)]\displaystyle-\frac{\sin 2(\phi_{M}-\delta_{3})-\sin 2(\phi_{S}-\delta_{3})}{2B_{3}^{2}}-\left.15\left(\frac{\phi_{R}-\phi_{S}}{B_{1}^{2}}+\frac{\phi_{M}-\phi_{R}}{B_{2}^{2}}-\frac{\phi_{M}-\phi_{S}}{B_{3}^{2}}\right)\right]
+ΛB1m3cotϕRcscϕRΛB1m3cotϕScscϕS.\displaystyle+\frac{\Lambda B_{1}m}{3}\cot\phi_{R}\csc\phi_{R}-\frac{\Lambda B_{1}m}{3}\cot\phi_{S}\csc\phi_{S}.
+ΛB2m3cot(ϕM+δ2)csc(ϕM+δ2)ΛB2m3cot(ϕR+δ2)csc(ϕR+δ2)\displaystyle+\frac{\Lambda B_{2}m}{3}\cot(\phi_{M}+\delta_{2})\csc(\phi_{M}+\delta_{2})-\frac{\Lambda B_{2}m}{3}\cot(\phi_{R}+\delta_{2})\csc(\phi_{R}+\delta_{2})
+ΛB3m3cot(ϕSδ3)csc(ϕSδ3)ΛB3m3cot(ϕMδ3)csc(ϕMδ3)\displaystyle+\frac{\Lambda B_{3}m}{3}\cot(\phi_{S}-\delta_{3})\csc(\phi_{S}-\delta_{3})-\frac{\Lambda B_{3}m}{3}\cot(\phi_{M}-\delta_{3})\csc(\phi_{M}-\delta_{3})
+𝒪(ε3).\displaystyle+{\cal O}(\varepsilon^{3}). (6.32)

In Eq. (6.32), the second to sixth lines correspond to the Schwarzschild-like part and the seventh to twelfth lines are order 𝒪(mΛ){\cal O}(m\Lambda) terms due to the cosmological constant Λ\Lambda. The contribution of the cosmological constant Λ\Lambda appears as 𝒪(Λm){\cal O}(\Lambda m) instead of 𝒪(Λ/m){\cal O}(\Lambda/m). Because ψpSdS\psi^{\rm SdS}_{p} can be expressed as Eq. (6.25), ψ1dS\psi^{\rm dS}_{1}, ψ2dS\psi^{\rm dS}_{2}, and ψ3dS\psi^{\rm dS}_{3} are completely eliminated from in αSdS\alpha_{\rm SdS}. Therefore, αSdS\alpha_{\rm SdS} does not include the terms described solely by the cosmological constant Λ\Lambda.

6.4 Integral Formula

We compute the total deflection angle αSdS\alpha_{\rm SdS} using the integral formula. As in the case of the angular formula, we represent the areal integral of the Gaussian curvature KSdSK^{\rm SdS} as

Σ3KSdS𝑑σSdS=Σ3KdS𝑑σdS+𝒪(m,m2,mΛ)terms,\displaystyle-\iint_{\Sigma^{3}}K^{\rm SdS}d\sigma^{\rm SdS}=-\iint_{\Sigma^{3}}K^{\rm dS}d\sigma^{\rm dS}+{\cal O}(m,m^{2},m\Lambda)~{}\mbox{terms}, (6.33)

where the areal integral of KdSK^{\rm dS} is given by Eq. (6.23).

The Gaussian curvature KSdSK^{\rm SdS} in the Schwarzschild–de Sitter spacetime becomes

KSdS=2mr3(13m2r+Λ6mr3Λr2)<0,\displaystyle K^{\rm SdS}=-\frac{2m}{r^{3}}\left(1-\frac{3m}{2r}+\frac{\Lambda}{6m}r^{3}-\Lambda r^{2}\right)<0, (6.34)

and the areal element dσSdSd\sigma^{\rm SdS} is given by

dσSdS=r(12mrΛ3r2)32drdϕ,\displaystyle d\sigma^{\rm SdS}=r\left(1-\frac{2m}{r}-\frac{\Lambda}{3}r^{2}\right)^{-\frac{3}{2}}drd\phi, (6.35)

Similar to the same way in which Eqs. (5.21) and (6.21) are integrated, we construct triangle Σ3\Sigma^{3}, which is bounded by three geodesics Γ1\Gamma_{1}, Γ2\Gamma_{2}, and Γ3\Gamma_{3}, and divide the triangle Σ3\Sigma^{3} into ΣRM3(ϕRϕϕM)\Sigma^{3}_{RM}(\phi_{R}\leq\phi\leq\phi_{M}) and ΣMS3(ϕMϕϕS)\Sigma^{3}_{MS}(\phi_{M}\leq\phi\leq\phi_{S}) (see Figure 3).

Before integrating the areal integral of the Gaussian curvature over the triangle Σ3\Sigma^{3}, we approximate the integrand of the areal integral up to the second order in ε=m/B\varepsilon=m/B as follows:

Σ3KSdS𝑑σSdS\displaystyle-\iint_{\Sigma^{3}}K^{\rm SdS}d\sigma^{\rm SdS} =Σ3Λ3r(1Λ3r2)32𝑑r𝑑ϕ\displaystyle=\iint_{\Sigma^{3}}\frac{\Lambda}{3}r\left(1-\frac{\Lambda}{3}r^{2}\right)^{-\frac{3}{2}}drd\phi
+Σ3{m[6Λr2(52Λr2)]3r2(1Λ3r2)52+m2[18Λr2(2110Λr2)]6r3(1Λ3r2)72}𝑑r𝑑ϕ\displaystyle+\iint_{\Sigma^{3}}\left\{\frac{m[6-\Lambda r^{2}(5-2\Lambda r^{2})]}{3r^{2}\left(1-\frac{\Lambda}{3}r^{2}\right)^{\frac{5}{2}}}+\frac{m^{2}[18-\Lambda r^{2}(21-10\Lambda r^{2})]}{6r^{3}\left(1-\frac{\Lambda}{3}r^{2}\right)^{\frac{7}{2}}}\right\}drd\phi
+𝒪(ε3).\displaystyle+{\cal O}(\varepsilon^{3}). (6.36)

Note that at this stage in Eq. (6.36), ε=ΛB2\varepsilon=\Lambda B^{2} is not treated as a small dimensionless expansion parameter.

The first term in the right-hand side of Eq. (6.36) has the same form as that of Eq. (6.23) in the de Sitter spacetime. However, the light trajectory rr is rpSdSr^{\rm SdS}_{p} instead of rpdSr^{\rm dS}_{p}. Let us compute the first term of Eq. (6.36). Substituting Eqs. (6.5), (6.6), and (6.7) into the first term of Eq. (6.36), integrating over rr, and remaining ε=m/B\varepsilon=m/B order terms, we have

Σ3Λ3r(1Λ3r2)32𝑑r𝑑ϕ\displaystyle\quad~{}\iint_{\Sigma^{3}}\frac{\Lambda}{3}r\left(1-\frac{\Lambda}{3}r^{2}\right)^{-\frac{3}{2}}drd\phi
=ϕRϕMr1SdSr2SdSΛ3r(1Λ3r2)32𝑑r𝑑ϕ+ϕMϕSr1SdSr3SdSΛ3r(1Λ3r2)32𝑑r𝑑ϕ\displaystyle=\int^{\phi_{M}}_{\phi_{R}}\int^{r^{\rm SdS}_{2}}_{r^{\rm SdS}_{1}}\frac{\Lambda}{3}r\left(1-\frac{\Lambda}{3}r^{2}\right)^{-\frac{3}{2}}drd\phi+\int^{\phi_{S}}_{\phi_{M}}\int^{r^{\rm SdS}_{3}}_{r^{\rm SdS}_{1}}\frac{\Lambda}{3}r\left(1-\frac{\Lambda}{3}r^{2}\right)^{-\frac{3}{2}}drd\phi
=ϕRϕM[(1Λ3r2)12]r1SdSr2SdS𝑑ϕ+ϕMϕS[(1Λ3r2)12]r1SdSr3SdS𝑑ϕ\displaystyle=\int^{\phi_{M}}_{\phi_{R}}\left[\left(1-\frac{\Lambda}{3}r^{2}\right)^{-\frac{1}{2}}\right]^{r^{\rm SdS}_{2}}_{r^{\rm SdS}_{1}}d\phi+\int^{\phi_{S}}_{\phi_{M}}\left[\left(1-\frac{\Lambda}{3}r^{2}\right)^{-\frac{1}{2}}\right]^{r^{\rm SdS}_{3}}_{r^{\rm SdS}_{1}}d\phi
=Σ3KdS𝑑σdS\displaystyle=-\iint_{\Sigma^{3}}K^{\rm dS}d\sigma^{\rm dS}
ϕRϕM{ΛB2m[3+cos2(ϕ+δ2)]csc3(ϕ+δ2)6(1Λ3B22csc2(ϕ+δ2))32ΛB1m(3+cos2ϕ)csc3ϕ6(1Λ3B12csc2ϕ)32}𝑑ϕ\displaystyle-\int^{\phi_{M}}_{\phi_{R}}\left\{\frac{\Lambda B_{2}m[3+\cos 2(\phi+\delta_{2})]\csc^{3}(\phi+\delta_{2})}{6\left(1-\frac{\Lambda}{3}B^{2}_{2}\csc^{2}(\phi+\delta_{2})\right)^{\frac{3}{2}}}-\frac{\Lambda B_{1}m(3+\cos 2\phi)\csc^{3}\phi}{6\left(1-\frac{\Lambda}{3}B^{2}_{1}\csc^{2}\phi\right)^{\frac{3}{2}}}\right\}d\phi
ϕMϕS{ΛB3m[3+cos2(ϕδ3)]csc3(ϕδ3)6(1Λ3B32csc2(ϕδ3))32ΛB1m(3+cos2ϕ)csc3ϕ6(1Λ3B12csc2ϕ)32}𝑑ϕ+𝒪(ε2).\displaystyle-\int^{\phi_{S}}_{\phi_{M}}\left\{\frac{\Lambda B_{3}m[3+\cos 2(\phi-\delta_{3})]\csc^{3}(\phi-\delta_{3})}{6\left(1-\frac{\Lambda}{3}B^{2}_{3}\csc^{2}(\phi-\delta_{3})\right)^{\frac{3}{2}}}-\frac{\Lambda B_{1}m(3+\cos 2\phi)\csc^{3}\phi}{6\left(1-\frac{\Lambda}{3}B^{2}_{1}\csc^{2}\phi\right)^{\frac{3}{2}}}\right\}d\phi+{\cal O}(\varepsilon^{2}). (6.37)

where ε=ΛB2\varepsilon=\Lambda B^{2} is also not still considered a small dimensionless expansion parameter. Next, expanding the fifth and sixth lines in Eq. (6.37) with respect to ε=ΛB2\varepsilon=\Lambda B^{2} and remaining 𝒪((m/B)ΛB2){\cal O}((m/B)\cdot\Lambda B^{2}) terms, we have

Σ3Λ3r(1Λ3r2)32𝑑r𝑑ϕ\displaystyle\quad~{}\iint_{\Sigma^{3}}\frac{\Lambda}{3}r\left(1-\frac{\Lambda}{3}r^{2}\right)^{-\frac{3}{2}}drd\phi
=Σ3KdS𝑑σdS\displaystyle=-\iint_{\Sigma^{3}}K^{\rm dS}d\sigma^{\rm dS}
+ΛB1m3cotϕRcscϕRΛB1m3cotϕScscϕS.\displaystyle+\frac{\Lambda B_{1}m}{3}\cot\phi_{R}\csc\phi_{R}-\frac{\Lambda B_{1}m}{3}\cot\phi_{S}\csc\phi_{S}.
+ΛB2m3cot(ϕM+δ2)csc(ϕM+δ2)ΛB2m3cot(ϕR+δ2)csc(ϕR+δ2)\displaystyle+\frac{\Lambda B_{2}m}{3}\cot(\phi_{M}+\delta_{2})\csc(\phi_{M}+\delta_{2})-\frac{\Lambda B_{2}m}{3}\cot(\phi_{R}+\delta_{2})\csc(\phi_{R}+\delta_{2})
+ΛB3m3cot(ϕSδ3)csc(ϕSδ3)ΛB3m3cot(ϕMδ3)csc(ϕMδ3)+𝒪(ε3),\displaystyle+\frac{\Lambda B_{3}m}{3}\cot(\phi_{S}-\delta_{3})\csc(\phi_{S}-\delta_{3})-\frac{\Lambda B_{3}m}{3}\cot(\phi_{M}-\delta_{3})\csc(\phi_{M}-\delta_{3})+{\cal O}(\varepsilon^{3}), (6.38)

where 𝒪(ε3)=𝒪((m/B)3,(m/B)2ΛB2,(m/B)(ΛB2)2){\cal O}(\varepsilon^{3})={\cal O}((m/B)^{3},(m/B)^{2}\cdot\Lambda B^{2},(m/B)\cdot(\Lambda B^{2})^{2}). Subsequently, we expand the integrand of the second line in Eq. (6.36) with respect to ε=ΛB2\varepsilon=\Lambda B^{2} remaining the order 𝒪((m/B)2,(m/B)ΛB2){\cal O}((m/B)^{2},(m/B)\cdot\Lambda B^{2}) terms. One has

Σ3{m[6Λr2(52Λr2)]3r2(1Λ3r2)52+m2[18Λr2(2110Λr2)]6r3(1Λ3r2)72}𝑑r𝑑ϕ\displaystyle\quad~{}\iint_{\Sigma^{3}}\left\{\frac{m[6-\Lambda r^{2}(5-2\Lambda r^{2})]}{3r^{2}\left(1-\frac{\Lambda}{3}r^{2}\right)^{\frac{5}{2}}}+\frac{m^{2}[18-\Lambda r^{2}(21-10\Lambda r^{2})]}{6r^{3}\left(1-\frac{\Lambda}{3}r^{2}\right)^{\frac{7}{2}}}\right\}drd\phi
=Σ3(2mr2+3m2r3)𝑑r𝑑ϕ+𝒪(ε3),\displaystyle=\iint_{\Sigma^{3}}\left(\frac{2m}{r^{2}}+\frac{3m^{2}}{r^{3}}\right)drd\phi+{\cal O}(\varepsilon^{3}), (6.39)

which is of the same form as Eq. (5.21) and 𝒪(ε3)=𝒪((m/B)3,(m/B)2ΛB2,(m/B)(ΛB2)2){\cal O}(\varepsilon^{3})={\cal O}((m/B)^{3},(m/B)^{2}\cdot\Lambda B^{2},(m/B)\cdot(\Lambda B^{2})^{2}). In this approximation, the order 𝒪(mΛ){\cal O}(m\Lambda) terms are eliminated from the integrand of Eq. (6.39). Integrating Eq. (6.39) gives

Σ3(2mr2+3m2r3)𝑑r𝑑ϕ\displaystyle\quad~{}\iint_{\Sigma^{3}}\left(\frac{2m}{r^{2}}+\frac{3m^{2}}{r^{3}}\right)drd\phi
=ϕRϕMr1SdSr2SdS(2mr2+3m2r3)𝑑r𝑑ϕ+ϕMϕSr1SdSr3SdS(2mr2+3m2r3)𝑑r𝑑ϕ\displaystyle=\int_{\phi_{R}}^{\phi_{M}}\int_{r^{\rm SdS}_{1}}^{r^{\rm SdS}_{2}}\left(\frac{2m}{r^{2}}+\frac{3m^{2}}{r^{3}}\right)drd\phi+\int_{\phi_{M}}^{\phi_{S}}\int_{r^{\rm SdS}_{1}}^{r^{\rm SdS}_{3}}\left(\frac{2m}{r^{2}}+\frac{3m^{2}}{r^{3}}\right)drd\phi
=2m[cosϕRcosϕSB1+cos(ϕM+δ2)cos(ϕR+δ2)B2+cos(ϕSδ3)cos(ϕMδ3)B3]\displaystyle=2m\left[\frac{\cos\phi_{R}-\cos\phi_{S}}{B_{1}}+\frac{\cos(\phi_{M}+\delta_{2})-\cos(\phi_{R}+\delta_{2})}{B_{2}}+\frac{\cos(\phi_{S}-\delta_{3})-\cos(\phi_{M}-\delta_{3})}{B_{3}}\right]
m24[sin2ϕRsin2ϕS2B12+sin2(ϕM+δ2)sin2(ϕR+δ2)2B22\displaystyle-\frac{m^{2}}{4}\left[\frac{\sin 2\phi_{R}-\sin 2\phi_{S}}{2B_{1}^{2}}+\frac{\sin 2(\phi_{M}+\delta_{2})-\sin 2(\phi_{R}+\delta_{2})}{2B_{2}^{2}}\right.
sin2(ϕMδ3)sin2(ϕSδ3)2B32ϕRϕSB12+ϕMϕRB22ϕMϕSB32)+𝒪(ε3),\displaystyle-\frac{\sin 2(\phi_{M}-\delta_{3})-\sin 2(\phi_{S}-\delta_{3})}{2B_{3}^{2}}-\left.\frac{\phi_{R}-\phi_{S}}{B_{1}^{2}}+\frac{\phi_{M}-\phi_{R}}{B_{2}^{2}}-\frac{\phi_{M}-\phi_{S}}{B_{3}^{2}}\right)+{\cal O}(\varepsilon^{3}), (6.40)

which corresponds to the Schwarzschild-like part given in Eqs. (5.17) and (5.21).

Substituting Eqs. (6.37), (6.38), and (6.40) into Eq. (6.36) and applying to Eq. (4.13), we have

αSdS\displaystyle\alpha_{\rm SdS} =|Σ3KSdS𝑑σSdSΣ3KdS𝑑σdS|\displaystyle=\left|\iint_{\Sigma^{3}}K^{\rm SdS}d\sigma^{\rm SdS}-\iint_{\Sigma^{3}}K^{\rm dS}d\sigma^{\rm dS}\right|
=2m[cosϕRcosϕSB1+cos(ϕM+δ2)cos(ϕR+δ2)B2+cos(ϕSδ3)cos(ϕMδ3)B3]\displaystyle=2m\left[\frac{\cos\phi_{R}-\cos\phi_{S}}{B_{1}}+\frac{\cos(\phi_{M}+\delta_{2})-\cos(\phi_{R}+\delta_{2})}{B_{2}}+\frac{\cos(\phi_{S}-\delta_{3})-\cos(\phi_{M}-\delta_{3})}{B_{3}}\right]
m24[sin2ϕRsin2ϕS2B12+sin2(ϕM+δ2)sin2(ϕR+δ2)2B22\displaystyle-\frac{m^{2}}{4}\left[\frac{\sin 2\phi_{R}-\sin 2\phi_{S}}{2B_{1}^{2}}+\frac{\sin 2(\phi_{M}+\delta_{2})-\sin 2(\phi_{R}+\delta_{2})}{2B_{2}^{2}}\right.
sin2(ϕMδ3)sin2(ϕSδ3)2B3215(ϕRϕSB12+ϕMϕRB22ϕMϕSB32)]\displaystyle-\frac{\sin 2(\phi_{M}-\delta_{3})-\sin 2(\phi_{S}-\delta_{3})}{2B_{3}^{2}}-\left.15\left(\frac{\phi_{R}-\phi_{S}}{B_{1}^{2}}+\frac{\phi_{M}-\phi_{R}}{B_{2}^{2}}-\frac{\phi_{M}-\phi_{S}}{B_{3}^{2}}\right)\right]
+ΛB1m3cotϕRcscϕRΛB1m3cotϕScscϕS.\displaystyle+\frac{\Lambda B_{1}m}{3}\cot\phi_{R}\csc\phi_{R}-\frac{\Lambda B_{1}m}{3}\cot\phi_{S}\csc\phi_{S}.
+ΛB2m3cot(ϕM+δ2)csc(ϕM+δ2)ΛB2m3cot(ϕR+δ2)csc(ϕR+δ2)\displaystyle+\frac{\Lambda B_{2}m}{3}\cot(\phi_{M}+\delta_{2})\csc(\phi_{M}+\delta_{2})-\frac{\Lambda B_{2}m}{3}\cot(\phi_{R}+\delta_{2})\csc(\phi_{R}+\delta_{2})
+ΛB3m3cot(ϕSδ3)csc(ϕSδ3)ΛB3m3cot(ϕMδ3)csc(ϕMδ3)\displaystyle+\frac{\Lambda B_{3}m}{3}\cot(\phi_{S}-\delta_{3})\csc(\phi_{S}-\delta_{3})-\frac{\Lambda B_{3}m}{3}\cot(\phi_{M}-\delta_{3})\csc(\phi_{M}-\delta_{3})
+𝒪(ε3).\displaystyle+{\cal O}(\varepsilon^{3}). (6.41)

Eq. (6.41) completely agrees with Eq. (6.32).

Before closing this section, it is noteworthy that in the case of asymptotically flat spacetime, we can say that there is no ambiguity about what is meant by the total deflection angle when the source and observer are sufficiently far away. However in the case of non-asymptotically flat spacetime, the total deflection angle depends on the choice of the triangle because different triangles in curved spacetime cause different sums of interior angles. Therefore, it is important to clarify how the triangle is constructed in a non-asymptotically flat spacetime.

6.5 Contribution of the Cosmological Constant and its Observability

Let us investigate how the cosmological constant Λ\Lambda contributes to the total deflection angle, and its observability.

Using Eqs. (6.32) and (6.41), we extract the part of the order 𝒪(Λm){\cal O}(\Lambda m) terms and set

αSdSΛ\displaystyle\alpha_{\rm SdS}^{\Lambda} =ΛB1m3cotϕRcscϕRΛB1m3cotϕScscϕS.\displaystyle=\frac{\Lambda B_{1}m}{3}\cot\phi_{R}\csc\phi_{R}-\frac{\Lambda B_{1}m}{3}\cot\phi_{S}\csc\phi_{S}.
+ΛB2m3cot(ϕM+δ2)csc(ϕM+δ2)ΛB2m3cot(ϕR+δ2)csc(ϕR+δ2)\displaystyle+\frac{\Lambda B_{2}m}{3}\cot(\phi_{M}+\delta_{2})\csc(\phi_{M}+\delta_{2})-\frac{\Lambda B_{2}m}{3}\cot(\phi_{R}+\delta_{2})\csc(\phi_{R}+\delta_{2})
+ΛB3m3cot(ϕSδ3)csc(ϕSδ3)ΛB3m3cot(ϕMδ3)csc(ϕMδ3).\displaystyle+\frac{\Lambda B_{3}m}{3}\cot(\phi_{S}-\delta_{3})\csc(\phi_{S}-\delta_{3})-\frac{\Lambda B_{3}m}{3}\cot(\phi_{M}-\delta_{3})\csc(\phi_{M}-\delta_{3}). (6.42)

We observe that the cosmological constant Λ\Lambda contributes to the total deflection angle, and that the leading terms have a form of 𝒪(Λm){\cal O}(\Lambda m) instead of 𝒪(Λ/m){\cal O}(\Lambda/m). Furthermore, the terms characterized only by the cosmological constant Λ\Lambda do not appear in the expression of the total deflection angle αSdS\alpha_{\rm SdS} (see Eqs. (6.21) and (6.23)). This is because angle ψSdS\psi^{\rm SdS} and the areal integral of Gaussian curvature KSdSK^{\rm SdS} can be expressed as Eqs. (6.25) and (6.33); therefore, the terms are completely eliminated, as seen in Eqs. (6.32) and (6.41).

Let us discuss the observability of the cosmological constant Λ\Lambda to the total deflection angle. We assume Λ1052m2\Lambda\approx 10^{-52}~{}{\rm m}^{-2} and consider the sun m=GM/c2m_{\odot}=GM_{\odot}/c^{2} and typical galaxy mgal1012GM/c2m_{\rm gal}\approx 10^{12}GM_{\odot}/c^{2} as the lens objects, where G=6.674×1011m3kg1s2G=6.674\times 10^{-11}~{}{\rm m^{3}\cdot kg^{-1}\cdot s^{-2}} denotes the Newtonian gravitational constant, c=3.0×108m2c=3.0\times 10^{8}~{}{\rm m}^{2} the speed of light in vacuum, and M=2.0×1030kgM_{\odot}=2.0\times 10^{30}~{}{\rm kg} the mass of the sun. Additionally, we employ the radius of sun and galaxy as the impact parameters b1bBR=6.960×108mb_{1}\approx b_{\odot}\approx B_{\odot}\approx R_{\odot}=6.960\times 10^{8}~{}{\rm m} and b1bgalBgalRgalaxy5.0×104ly5×1020mb_{1}\approx b_{\rm gal}\approx B_{\rm gal}\approx R_{\rm galaxy}\approx 5.0\times 10^{4}~{}{\rm ly}\approx 5\times 10^{20}~{}{\rm m}, respectively.

When the lens object is the sun, we observe

4mB8.5×106,15πm24B25.3×1011,ΛBm33.3×1041,\displaystyle\frac{4m_{\odot}}{B_{\odot}}\approx 8.5\times 10^{-6},\quad\frac{15\pi m^{2}_{\odot}}{4B_{\odot}^{2}}\approx 5.3\times 10^{-11},\quad\frac{\Lambda B_{\odot}m_{\odot}}{3}\approx 3.3\times 10^{-41}, (6.43)

where the unit is in rad (radian). The order of the contribution of the cosmological constant αSdSΛ\alpha_{\rm SdS}^{\Lambda} is 𝒪(1041)rad{\cal O}(10^{-41})~{}~{}{\rm rad}. Even if we consider the cotϕcscϕ\cot\phi\csc\phi term, αSdSΛ\alpha_{\rm SdS}^{\Lambda} is only 𝒪(1036){\cal O}(10^{-36}) at most where we assumed that observer RR and source SS are located at the position of Earth and its opposition, i.e., ϕR=arcsinB/(1.5×1011)\phi_{R}=\arcsin B_{\odot}/(1.5\times 10^{11}), and ϕS=πarcsinB/(1.5×1011)\phi_{S}=\pi-\arcsin B_{\odot}/(1.5\times 10^{11}), respectively.

However, when the lens object is the galaxy, we observe

4mgalBgal1.2×105,15πmgal24Bgal21.0×1010,ΛBgalmgal32.5×1017.\displaystyle\frac{4m_{\rm gal}}{B_{\rm gal}}\approx 1.2\times 10^{-5},\quad\frac{15\pi m^{2}_{\rm gal}}{4B_{\rm gal}^{2}}\approx 1.0\times 10^{-10},\quad\frac{\Lambda B_{\rm gal}m_{\rm gal}}{3}\approx 2.5\times 10^{-17}. (6.44)

Accordingly, αSdSΛ\alpha_{\rm SdS}^{\Lambda} is αSdSΛ𝒪(1017)\alpha_{\rm SdS}^{\Lambda}\approx{\cal O}(10^{-17}) and mostly three orders of magnitude smaller than the sensitivity of the angle observed in planned space missions (such as LATOR), 0.01picorad=1014rad0.01~{}{\rm picorad}=10^{-14}~{}{\rm rad} (see Figure 3 in [26]).

However, because of the cotϕcscϕ\cot\phi\csc\phi term, αSdSΛ\alpha_{\rm SdS}^{\Lambda} rapidly increases when source SS and receiver RR reach the de Sitter horizon rrΛ=3/Λ1.73×1026mr\rightarrow r_{\Lambda}=\sqrt{3/\Lambda}\approx 1.73\times 10^{26}~{}{\rm m}, where we estimated the angular coordinate of the de Sitter horizon as ϕΛ=arcsin(Bgal/rΛ)3.0×106\phi_{\Lambda}=\arcsin(B_{\rm gal}/r_{\Lambda})\approx 3.0\times 10^{-6}, and suppose triangle Σ3\Sigma^{3} to be symmetrical with respect to ϕ=π/2\phi=\pi/2 (see Figures 4 and 5). Let source SS and receiver RR located in a symmetrical position near the de Sitter horizon with ϕS=ϕΛ\phi_{S}=\phi_{\Lambda} and ϕR=πϕΛ\phi_{R}=\pi-\phi_{\Lambda}. Furthermore, because of the same reason as that in Eq. (5.23), SS and RR approach the de Sitter horizon, and b2=b3b_{2}=b_{3} also approaches the de Sitter horizon, i.e., b2=b3rΛb_{2}=b_{3}\rightarrow r_{\Lambda}.

However, it is reasonable to set

cot(π2δ2)csc(π2δ2)\displaystyle\cot\left(\frac{\pi}{2}-\delta_{2}\right)\csc\left(\frac{\pi}{2}-\delta_{2}\right) =cot(π2+δ2)csc(π2+δ2)𝒪(1)\displaystyle=-\cot\left(\frac{\pi}{2}+\delta_{2}\right)\csc\left(\frac{\pi}{2}+\delta_{2}\right)\simeq{\cal O}(1) (6.45)
cot(ϕΛ+δ2)csc(ϕΛ+δ2)\displaystyle\cot(\phi_{\Lambda}+\delta_{2})\csc(\phi_{\Lambda}+\delta_{2}) cotδ1cscδ2𝒪(10)𝒪(102)\displaystyle\simeq\cot\delta_{1}\csc\delta_{2}\simeq{\cal O}(10)\sim{\cal O}(10^{2}) (6.46)
cot(πϕΛδ2)csc(πϕΛδ2)\displaystyle\cot(\pi-\phi_{\Lambda}-\delta_{2})\csc(\pi-\phi_{\Lambda}-\delta_{2}) cot(πδ2)csc(πδ2)𝒪(10)𝒪(102),\displaystyle\simeq\cot(\pi-\delta_{2})\csc(\pi-\delta_{2})\simeq{\cal O}(10)\sim{\cal O}(10^{2}), (6.47)

where we assumed 5δ2305\lesssim\delta_{2}\lesssim 30 degree. Then, we can estimate as

Λb2mgal3cot(π2+δ2)csc(π2+δ2)=Λb2mgal3cot(π2δ2)csc(π2δ2)1011,\displaystyle\frac{\Lambda b_{2}m_{\rm gal}}{3}\cot\left(\frac{\pi}{2}+\delta_{2}\right)\csc\left(\frac{\pi}{2}+\delta_{2}\right)=-\frac{\Lambda b_{2}m_{\rm gal}}{3}\cot\left(\frac{\pi}{2}-\delta_{2}\right)\csc\left(\frac{\pi}{2}-\delta_{2}\right)\simeq 10^{-11}, (6.48)
Λb2mgal3cotδ2cscδ2=Λb2mgal3cot(πδ2)csc(πδ2)1010109.\displaystyle\frac{\Lambda b_{2}m_{\rm gal}}{3}\cot\delta_{2}\csc\delta_{2}=-\frac{\Lambda b_{2}m_{\rm gal}}{3}\cot(\pi-\delta_{2})\csc(\pi-\delta_{2})\simeq 10^{-10}\sim 10^{-9}. (6.49)

Accordingly, αSdSΛ\alpha^{\Lambda}_{\rm SdS} can be evaluated as

αSdSΛ5.9×106rad,\displaystyle\alpha^{\Lambda}_{\rm SdS}\approx-5.9\times 10^{-6}~{}{\rm rad}, (6.50)

which is almost half the value of the Schwarzschild part. Hence, if both source SS and receiver RR are located near the de Sitter horizon, we may be able to detect the contribution of the cosmological constant Λ\Lambda to the total deflection angle.

7 Summary and Conclusions

Assuming a static and spherically symmetric spacetime, we proposed a new concept of the total deflection angle of a light ray in curved spacetime by means of the optical geometry which is considered as the Riemannian geometry experienced by the light ray. The concept is defined by the difference between the sum of internal angles of two triangles; one of them lies on curved spacetime distorted by a gravitating body and the other on the background spacetime. Our new definition of the total deflection angle is geometrically and intuitively clear. The triangle used to define the total deflection angle was realized by setting three laser-beam baselines (i.e., three null geodesics) in the space, inspired from planned space missions, including, LATOR, ASTROD-GW, and LISA. Accordingly, our new total deflection angle can be calculated by measuring the difference between the sum of the internal angles of both the triangles. Two formulas were presented to calculate the total deflection angle, in accordance with the Gauss–Bonnet theorem. It was shown that in the case of the Schwarzschild spacetime, the expression of the total deflection angle reduced to Epstein–Shapiro’s formula when the source of the light ray, SS, and observer RR were located in an asymptotically flat region. Additionally, in the case of the Schwarzschild–de Sitter spacetime, the total deflection angle was represented by the Schwarzschild-like part and the coupling terms of the central mass mm and cosmological constant Λ\Lambda as the form of 𝒪(Λm){\cal O}(\Lambda m) instead of 𝒪(Λ/m){\cal O}(\Lambda/m). Furthermore, the expression for the total deflection angle did not include the terms characterized solely by the cosmological constant Λ\Lambda; this is obvious from the fact that the angle ψpSdS\psi^{\rm SdS}_{p} and the area integrals of the Gaussian curvature KSdSK^{\rm SdS} in the Schwarzschild–de Sitter spacetime can be expressed using those in the de Sitter spacetime, ψpdS\psi^{\rm dS}_{p} and KdSK^{\rm dS} as described in Eqs (6.29), (6.30), (6.31), and (6.38), see also (6.25) and (6.33).

When the lens object was the sun, the magnitude of the contribution of the cosmological constant to the total deflection angle was αSdSΛ𝒪(1036)\alpha_{\rm SdS}^{\Lambda}\simeq{\cal O}(10^{-36}). Accordingly, it was extremely difficult to detect the contribution. However, if the galaxy was the lens object, the contribution of the cosmological constant was approximately αSdSΛ𝒪(1017)\alpha_{\rm SdS}^{\Lambda}\simeq{\cal O}(10^{-17}), which is mostly three orders of magnitude smaller than the sensitivity 0.01picorad=1014rad0.01~{}{\rm picorad}=10^{-14}~{}{\rm rad} observed in planned space missions such as LATOR. However, when the source of the light ray, SS, and the observer RR reached the de Sitter horizon, αSdSΛ\alpha_{\rm SdS}^{\Lambda} became αSdSΛ5.9×106rad\alpha_{\rm SdS}^{\Lambda}\simeq-5.9\times 10^{-6}~{}{\rm rad}, which is almost half of the Schwarzschild part. Therefore, if we can observe gravitational lensing of distant galaxies such as the recently reported RXCJ0600-z6 [38] which is located almost at the de Sitter horizon, 12.9billionly1.22×1026m(z=6.0719)12.9~{}{\rm billion\ ly}\simeq 1.22\times 10^{26}~{}{\rm m}~{}(z=6.0719), it may be possible to detect the contribution of the cosmological constant to the bending of light using the gravitational lensing effect.

Especially for observations in the solar system, the clocks on board the three satellites need to be synchronized accurately. Then the problem of synchronization of the clocks must be discussed as a future problem when our results are applied to actual measurements.

Gravitational lensing may pave the way to solving the cosmological constant problem. However, it is currently difficult to use our formulation directly for cosmological constant/dark energy exploration using gravitational lensing, and further extensions and improvements are required. This is because first we must arrange the polygons such that the singularity of the central object is successfully incorporated into the gravitational lensing equation. And second, in our formulation we need to place the observer at the three vertices of the triangle. However, in the actual gravitational lensing effect, the observer can only be placed at one of the three vertices of the triangle 444 Nevertheless, although there are some problems in practical applications, we think that our results are useful because we can show how the formula of the total deflection angle of light ray is affected by the cosmological constant without ambiguity. . Therefore, we have also to resolve these problems in the future.

Acknowledgments

We would like to acknowledge anonymous referee for reading our paper carefully and for giving fruitful comments and suggestions, which significantly improved the quality of the paper. We also appreciate H. Asada and K. Takizawa for fruitful discussions and comments. This work was partially supported by Dean’s Grant for Specified Incentive Research, College of Engineering, Nihon University.

References

  • [1] Riess, A. G.; Filippenko, A. V.; Challis, P.; Clocchiatti, A.; Diercks, A.; Garnavich, P. M.; Gilliland, R. L.; Hogan, C. J.; Jha, S.; Kirshner, R. P.; Leibundgut, B.; Phillips, M. M.; Reiss, D.; Schmidt, B. P.; Schommer, R. A.; Smith, R. C.; Spyromilio, J.; Stubbs, C.; Suntzeff, N. B.; Tonry, J., Astron. J. 1998, 116, 1009-1038.
  • [2] Schmidt, B. P.; Suntzeff, N. B.; Phillips, M. M.; Schommer, R. A.; Clocchiatti, A.; Kirshner, R. P.; Garnavich, P.; Challis, P.; Leibundgut, B.; Spyromilio, J.; Riess, A. G.; Filippenko, A. V.; Hamuy, M.; Smith, R. C.; Hogan, C.; Stubbs, C.; Diercks, A.; Reiss, D.; Gilliland, R.; Tonry, J.; Maza, J.; Dressler, A.; Walsh, J.; Ciardullo, R., Astrophys. J., 1998, 507, 46-63.
  • [3] Perlmutter, S.; Aldering, G.; Goldhaber, G.; Knop, R. A.; Nugent, P.; Castro, P. G.; Deustua, S.; Fabbro, S.; Goobar, A.; Groom, D. E.; Hook, I. M.; Kim, A. G.; Kim, M. Y.; Lee, J. C.; Nunes, N. J.; Pain, R.; Pennypacker, C. R.; Quimby, R.; Lidman, C.; Ellis, R. S.; Irwin, M.; McMahon, R. G.; Ruiz-Lapuente, P.; Walton, N.; Schaefer, B.; Boyle, B. J.; Filippenko, A. V.; Matheson, T.; Fruchter, A. S.; Panagia, N.; Newberg, H. J. M.; Couch, W. J.; Astrophys. J., 1999, 517, 565-586.
  • [4] Schneider, P.; Ehlers, J.; Falco, E. E.: Gravitational Lenses, Springer Verlag, Berlin, Heidelberg, New York (1999).
  • [5] Schneider, P.; Kochanek, C.; Wambsganss, J.: Gravitational Lensing: Strong, Weak and Micro, Springer, Berlin, Heidelberg, New York (2006).
  • [6] Islam, J. N., Phys. Lett. A, 97, 239-241 (1983).
  • [7] Rindler, W.; Ishak, M., Phys. Rev. D, 76, id. 043006 (2007).
  • [8] Ishak, M.; Rindler, W., Gen. Rel. Grav., 42, 2247 (2010).
  • [9] Lake, K., Phys. Rev. D, 65, id. 087301 (2002).
  • [10] Park, M., Phys. Rev. D, 78, id. 023014 (2008).
  • [11] Khriplovich, I. B.; Pomeransky, A. A., Int. J. Mod. Phys. D, 17, 2255 (2008).
  • [12] Simpson, F.; Peacock, J. A.; Heavens, A. F., MNRAS, 402, 2009 (2010).
  • [13] Bhadra, A.; Biswas, S.; K. Sarkar, Phys. Rev. D, 82, id.063003 (2010).
  • [14] Miraghaei, H.; Nouri-Zonoz, M.; Gen. Rel. Grav., 42, 2947 (2010).
  • [15] Biressa, T.; de Freitas Pacheco, J. A., Gen. Rel. Grav., 43, 2649 (2011).
  • [16] Arakida, H.; Kasai, M., Phys. Rev. D, 85, id.023006 (2012).
  • [17] Hammad, F. Mod. Phys. Lett. A, 28, 1350181 (2013).
  • [18] Lebedev, D.; Lake, K., arXiv:1308.4931 (2013).
  • [19] Batic, D.; Nelson, S.; Nowakowski, M., Phys. Rev. D, 91, id.104015 (2015).
  • [20] Arakida, H., Universe, 2, 5 (2016).
  • [21] Gibbons, G. W.; Werner, M. C., Class. Quant. Grav., 25, id. 235009 (2008).
  • [22] Ishihara, A.; Suzuki, Y.; Ono, T.; Kitamura, T., Asada, H., Phys. Rev. D, 94, id.084015 (2016).
  • [23] Ishihara, A.; Suzuki, Y.; Ono, T.; Asada, H., Phys. Rev. D, 95, id.044017 (2017).
  • [24] Arakida, H., Gen. Relativ. Gravit., 50, 48 (2018).
  • [25] Takizawa, K.; Ono, T; Asada, H., Phys. Rev. D, 101, id.104032 (2020).
  • [26] Turyshev, S. G.; Shao, M.; Nordtvedt, K. L.; Dittus, H.; Laemmerzahl, C.; Theil, S.; Salomon, C.; Reynaud, S.; Damour, T.; Johann, U.; Bouyer, P.; Touboul, P.; Foulon, B.; Bertolami, O.; Páramos, J., Exp Astron., 27, 27 (2009).
  • [27] Ni, W. T.; Men, J. R.; Mei, X. H. et al. 2009 Proc. Sixth Deep Space Exploration Technology Symposium, 122 (2009).
  • [28] Amaro-Seoane, P. et al., arXiv:1702.00786 (2017).
  • [29] Abramowicz, M. A.; Carter, B.; Lasota, J. P., Gen. Rel. Grav., 20, 1173 (1988).
  • [30] Carroll, S., Spacetime and Geometry: An Introduction to General Relativity, (Addison Wesley, San Francisco, 2004).
  • [31] Klingenberg, W., A Course in Differential Geometry, (Springer Verlag, New York, 1978).
  • [32] Kreyszig, E., Differential Geometry, (Dover Publications, New York, 1991).
  • [33] do Carmo, M. P., Differential Geometry of Curves and Surfaces, 2nd ed., (Dover Publications, Mineola, New York, 2016).
  • [34] Epstein, R.; Shapiro, I. I., Phys. Rev. D, 22, 2947 (1980).
  • [35] Kottler, F., Annalen. Phys. 361, 401 (1918).
  • [36] Jeffrey, A.; Zwillinger, D. (Eds), Table of Integrals, Series, and Products, 7th ed., (Academic Press, 2007).
  • [37] Rindler, W., Relativity: Special, General, and Cosmological, 2nd ed., (Oxford University Press, New York, 2006).
  • [38] Fujimoto, S.; Oguri, M.; Brammer, G.; Yoshimura, Y.; Laporte, N.; González-López, J.; Caminha, G. B.; Kohno, K.; Zitrin, A.; Richard, J.; Ouchi, M.; Bauer, F. E.; Smail, I.; Hatsukade, B.; Ono, Y.; Kokorev, V.; Umehata, H.; Schaerer, D.; Knudsen, K.; Sun, Fengwu Magdis, G.; Valentino, F.; Ao, Y.; Toft, S.; Dessauges-Zavadsky, M.; Shimasaku, K.; Caputi, K.; Kusakabe, H.; Morokuma-Matsui, K.; Shotaro, K.; Egami, E.; Lee, Minju M.; Rawle, T.; Espada, D.; Astrophys. J., 911, id.99 (2021).